You are on page 1of 48

A special type of complex.

here called an ion-pair,


is an association of equally clharged units to form a
neutral species:

Chemical analyses of water generally report the


total quantity of a particular element or ion without
indicating its actual form in solution. For usein chemical
thermodynamic calculations, the concentrations of participating reactants or products must be identified as
specific solute species.
Complex ions are solute speciesmade up from two
or more single ions of opposite charge. For example,
ferric iron in solution may hydrolyze to form a hydroxide
complex:

Ca2++S042f CaS04(aq).
In this complex, the Cazt and S04- components are
believed to be separatedby one:or more layers of intervening water molecules. Some authors have used the
term ion pair in a less rigorous context.
Polynuclear complexes are defined by Baesand
Mesmer (1976, p. 2) ascomplexes containing more than
one cation. This is a broader definition than the one
implied above for polymeric species. The term polynuclear will not be used in this book.

Fe3++Hz0= FeOH++H+.
The Fe3 and OH- in this unit are chemically bound to
each other. If the concentration of ferric iron is high
enough, the formation of complex species may also
involve association of monomeric units to give dimers

SIGNIFICANCE
OF PROPERTIES
CONSTITUENTS
REPORTED IN
WATER ANALYSES

2FeOH - [Fe(OH),Fe14+,
and hydrolysis and polymerization may continue, giving
a final product approaching the composition Fe(OH)30
that contains many individual Fe3+and OH- ions linked
together.
Table 10. Chemical

analysis

The propertiesand constituents that are determined


in water analysesare discussedindividually in the following sections.For most constituents, the subjects considered are the form of dissolved species,solubility and

of a water

sample

expressed

[Sample from flowing well, 488 ft deep. Water from the Lance Formation.
Sheridan County, Wyo. Collected August 3, 19461
mg/L

Constituent

or rwm

.._..._._
7.9
Iron (Fe) ._______________.17
Calcium (Ca) .._.___ 37
Magnesium (Mg) ____ 24
Sodium (Na) ._._______
611
Potassium (K) ________
>
Bicarbonate
(HCOs) .._..._....____.__429
Sulfate (S04) _______.
1,010
Chloride (Cl) _____._. 82
Fluoride (F) _________. .6
Nitrate (NOs) ________ .O
Boron (B) ___._........_
.2
Dissolved solids:
Calculated _________.__
1,980
Hardness as CaCOs:
191
Total ___________.........
Noncarbonate ______ 0
Specific conduct2,880
ante (micromhos at
2YC).
7.3
pH _____________.

meq/L
or epm

Silica (SiO2)

1.85
1.97

mM/L

NW l/4

Cravimetric
percent

0.131
.003
.925
.985

Study

and Interpretation

of the Chemical

set 30, T. 57 N., R. 85 W.,

Percentage
of epm

Grains per
U.S. gallon

0.40
.Ol
1.87
1.21

6.1
6.5

0.46
.Ol
2.16
1.40

30.80

87.4

35.69

26.58

7.03
21.03
2.31
.03
.oo

7.03
10.52
2.31
,032
,000
.019

10.63
50.90
4.14
.03
.oo
.Ol

23.1
69.2
7.6
.I
.O
_...............

25.06
59.00
4.79
.04
.oo
.Ol

60.80

48.535

100.00

200.00

115.65

7.3

1.91
.oo
2,880

11.16
_________.____............................0
2,880
2,880
2,880

7.3

7.3

As carbonate (COa).

58

in six ways

26.58

2,880

Characteristics

AND

of Natural

Water

7.3

7.3

other concentration controls in relation to other dissolved


substances,the source of the material, the range of concentration that may be expected, and factors that may
influence the accuracy or precision of analytical determinations. Appropriate environmental influences are
considered,and chemical analysesare used to illustrate
the discussions. Literature references are given to aid
readers who may be interested in learning more about
the topic.
The chemistry of some solutes in natural water is
reasonably well understood. For others, much remains
to be learned. New research results will tend to make
many statementsmade here subject to rather rapid obsolescence.This is one reasonfor keeping the discussionsof
individual constituents generalizedand rather brief.
Nature

of the Dissolved

State

The discussions so far have presupposedthat the


dissolved state was well enough understood that it did
not need to be defined. An exact definition of the term,
however,doesposesomedifficulties. An aqueoussolution
can be defined as a uniformly homogeneousmixture of
water and solute units that constitutes a single phase.
Each of the dissolved solute units is surrounded by
solvent molecules and has no direct contact with other
like particles except as permitted by migration through
the solution. The particles may carry an electrical charge,
or they may be electrically neutral. There can be no
doubt that individual ions, or pairs of ions, or even
complexes made up of several ions, are in the dissolved
state. If the particles are large aggregatesthat will settle
out of the solvent by gravity, they certainly can no longer
be said to be in the dissolved state.
Betweentheseextremeslie particles of widely varying sizesthat may, under favorableconditions, be retained
in suspensionindefinitely. These particles generally are
consideredcolloidal, and the smaller colloidal particles
merge into the dissolved state. Some textbooks suggest
size ranges;for example, Glasstoneand Lewis (1960, p.
571) stated that colloidal particles range in diameter
from about 5 nanometers (nm) up to about 0.2 micrometer (pm). A micrometer is lo3 nanometers, or lo4
angstrom units (A). Therefore by this definition, the
lower limit of diameter of colloidal particles is 50 A. A
considerable number of organic compounds that may
occur in natural water, such asthosethat impart a brown
color, havemoleculesthat are more than 50 A in diameter;
hence, by this definition, they can occur only in the
colloidal state, even though all particles may be single
molecules. Inorganic polymeric ions may also attain
collodial size. For example, aluminum hydroxide polymers form charged units containing hundreds of aluminum ions linked together by hydroxide ions (Smith and
Hem, 1972, p. D38) that would have diameters between
Significance

50 and 100 A. Data presentedby Smith and Hem (1972,


p. D38) indicate that there is a change in chemical
behavior of these polymers when they attain this size,
which suggeststhat a transition between dissolved and
particulate statesoccurs at that point.
Natural surface waters and ground waters carry
both dissolved and suspendedparticles. The amounts of
the latter presentin ground water before it is brought to
the land surface generally are small. But in river water,
the concentration of suspendedmaterial may be large,
and at a high stage of flow in many streams it greatly
exceedsthe concentration of dissolved solids. This suspended material posessomething of a dilemma for the
water chemist. In the usual water-chemistry study, the
suspendedsolids are removed before analysis either by
filtration or by settling. Both proceduresremove particles
more than a few tenths of a micrometer in diameter, that
is, particles large enough to produce substantial light
scattering. The suspendedsolids are then discarded, and
the clear solution is analyzed. Some kinds of suspended
material, however, may have an important bearing on
the sanitary condition of the water. For example,analyses
for the biochemical oxygen demand or for bacterial
counts must be made using unfiltered samples.For many
of the river-sampling sites used by the U.S. Geological
Survey, the quantity of suspended load of streams is
determined, but, except for separation into size ranges,
very little has been done to determine the chemical
composition and related propertiesof suspendedloads of
streams. Water users can justifiably say they are not
interested in the composition of material they would
filter out of the water before they used it in any event;
hence, for many practical water-use applications the
practice of filtering a sample and analyzing only the
filtrate is entirely correct.
Characterization
of Suspended
Particulate
Material

Particle-size distributions of suspended sediment


are determined by various procedures, including wet
sieving and analyzing differences in settling rate (Guy,
1969). Instrumental proceduresalso have become available (Ritker and Helley, 1969). Another physical property that has considerable significance is the effective
surface area of sediment per unit weight. Surface area
measurementsof sediment fractions would have a more
direct bearing on the physical-chemical behavior of the
material than would the particle-size measurements.
A generalizedterminology usedin many U.S. Geological Survey reports calls sediment having particle
diametersranging from 0.24 I.trnto 4 pm clay. Material
in the range from 4 to 62 pm is termed silt, and that in
the rangefrom 62 pm to 2.0 mm is termed sand. In this
context, clay does not have mineralogic significance.
of Properties

and Constituents

Reported

in Water

Analyses

59

fresh or saline water.


Technologies-for evaluating the adsorbed or readily
available fractions of solutes held by stream sediment
have been suggested by Gibbs (1973) Malo (1977), and
many others. However, methods that will produce reliable,
complete, and reproducible results at what would generally be considered acceptable cost, for routine datacollection purposes, do not presently exist.
In the first place, separation of the solution phase
from the solids is required. It should be evident to the
most uninitiated student of water chemistry that the
chemical and physical behavior of the adsorbed material
will differ from that of the dissolved material both in the
steam itself and during any treatment process. Determinations of inorganic constituents on unfiltered samples,
which is practiced by some la,boratories that analyze
river water routinely, give results that can be interpreted
only in a gross, qualitative way. Unfortunately, also,
some of the data obtained by this procedure have been
stored in data banks or have been published without any
clear indication that they represent unfiltered samples.
Separation of the phases is routinely done by filtration in the field through membrane filters having pores
0.45 pm in diameter. This size opening has come to be
generally adopted as defining clissolved, as opposed to
particulate, material. Most of the time, however, samples having significant sediment contents tend to plug
many of the filter pores and effectively decrease the pore
diameters to some value less than 0.45 pm. Observations
of the effectiveness of filtration in removing fine particles
of aluminum, iron, manganese, and titanium oxides and
hydroxides from river water were reported by Kennedy
and others (1974). These investigators found that filter
membranes having pores 0.10 pm in diameter were
considerably more effective for removing these materials
than were membranes having larger pore diameters.
After filtration, the preferable procedure is to analyze
both phases. Part of the filtrate is usually stabilized by
acidification, which generally dissolves any colloidal
material that might be present and minimizes adsorption
of cations on container walls. Major constituents are
determined using an unacidified fraction of the filtrate,
because some constituents would be altered or destroyed
by acid treatment. Analysis of the filtrate is straightforward. Characterization of the residue on the filter is
much more difficult. The amount of original sample
filtered needs to be known so that results may be properly
assigned to a specific volume of water. The total weight
of residue must be determined. Various types of treatment
may then be used, either serially or on separate aliquots,
to recover potentially soluble material. The treatments
that have been used to recover the material range from
simple to very complex, from extractions with neutral
solutions of ammonium salts to displace adsorbed ions,
through mild reducing agents.. dilute acids or bases,

Chemical composition of sediment can be determined by methods of rock analysis in which all constituents are rendered soluble by strong acid treatment or
fusion. The composition may be expressed in mineralogic
terms, generally by identifying species present by optical
methods or by X-ray diffraction. Besides the various
aluminosilicate minerals, metal oxide, hydroxide, and
organic materials commonly are present, especially in
the finer size classes. The oxide, hydroxide, and organic
material also may be present on surfaces of clay particles
or other silicate mineral particles in all size ranges. Solute
cations are adsorbed on the particle surfaces as well, in
proportions related to their relative dissolved concentrations and physicochemical properties of the ions and
surfaces.
Physicochemical properties of sediment material
that are important in evaluating their behavior toward
solutes include the distribution, sign, and intensity of
surface electrical charge and the way such charges interact
with solute ions. The determinations that can be made to
evaluate these properties include charge-site distribution
(usually in equivalents or moles of positive or negative
charge per unit area), zero point of charge (zpc), cationexchange capacity (in equivalents per unit weight), and
distribution or selectivity coefficients.
The zpc is defined as the pH at which the surface
charge is neutral or effectively zero. At lower pHs the
charge would be positive, and at higher ones, negative.
The other measurements were described earlier in the
section titled Reactions at Interfaces.
Solute-sediment interactions obviously are fraught
with complexity. Many attempts to devise methods and
techniques for practical evaluation of the effects have
been made, but much more work will be required before
a good understanding of processes at particle surfaces
can be reached.
Gibbs (1973) presented some preliminary data evaluating the relative importance of some solute-sediment
interactions in the Amazon and Yukon Rivers. Kennedy
(1965) studied the mineralogy and exchange capacity of
sediments from 21 U.S. streams and estimated that in
many streams, when flow and sediment concentrations
were high, the ratio of total adsorbed to total dissolved
cations exceeded 0.1 and in some streams was greater
than 1.0. The streams studied represented a wide range
of geologic and hydrologic condtions. A more recent
study of the Amazon by Sayles and Mangelsdorf (1979)
also showed that at flood stages the ratio exceeded 0.1,
but these writers believed the adsorbed fraction was
insignificant on an average annual basis because the
Amazon usually carries low sediment concentrations.
On reaching the ocean, many of the adsorbed ions are
probably exchanged for sodium. The adsorbed ions also
may be exchanged for solute ions in the stream water if
the composition of the water changes owing to inflows of
60

Study

and

interpretation

of the

Chemical

Characteristics

of

Natural

Water

oxidizing acids, and, finally, dissolution in hydrofluoric


acid mixtures or through alkali fusion procedures used in
rock analysis.
Owing to the complexity of some of these analysis
schemes and the difficulty in carrying them out or in
interpreting the results, simpler analytical procedures
have been used in the U.S. Geological Survey water
laboratories. A determination of total recoverable inorganic constituents is made using a specific volume of
unfiltered sample that is acidified with hydrochloric acid
to a molarity near 0.3 in H and is held at a temperature
near the boiling point for 30 minutes. The sample is then
filtered and the required analytical determinations are
made on the filtrate (Skougstad and others, 1979).
Data obtained in this way must be supplemented by
comparable analyses of filtered fractions of the same
sample. The amount of total solute that was associated
with the solid phase can then be estimated.
The bulk composition of sediments carried to the
ocean by many large rivers was determined by Martin
and Meybeck (1979). They also estimated proportions
of the various elements transported in suspended as
opposed to dissolved forms. According to these authors,
the ratio of dissolved to total quantity transported exceeds
0.5 only for bromine, iodine, sulfur, chlorine, calcium,
sodium. and strontium.

Hydrogen-Ion

Activity

(pH)

The effective concentration (activity) of hydrogen


ions could be expressed in the same kinds of units as
other dissolved species, but H concentrations in milligrams per liter or moles per liter are very low for water
solutions that are not strongly acid. The activity of hydrogen ions can be expressed most conveniently in logarithmic units, and the abbreviation pH represents the
negative base-10 log of the hydrogen-ion activity in
moles per liter.
Theoretical concepts and various practical aspects
of pH determination have been discussed at length in the
literature. Bates (1973) has prepared an excellent and
authoritative summary of these topics. The notation
pH is now generally taken to mean hydrogen-ion
activity rather than concentration, although the distinction
between these concepts was not understood at the time
Sorensen proposed the use of the pH notation in 1909.
Throughout this discussion the term hydrogen ion is
used with the reservation that such species exist only in
hydrated form in aqueous solution.
Even when no other solutes are present, a few of the
Hz0 molecules in liquid water will be broken up into H
Significance

and OH- ions. This process of dissociation is a chemical


equilibrium that may be written
H,O(l)=H++OH-.
In mass-law form, the equilibrium
the equation

can be expressed as

By convention, the activity of the liquid water is taken to


be unity in this very dilute solution, and the constant K,
is then equal to the product of the activities of H and
OH-. This ion-activity product for water at 25C is, in
exponential terms, lo-l4 Ooo(Covington and others, 1966).
The two-place log of K.. is -14.00. At neutrality, by
definition [H]=[OH-] and therefore pH=7.00.
At higher temperatures, K,. increases and the neutral
value of pH becomes smaller; the value for 30C given
by Covington and others (1966) is lo-r3 837.Neutral pH
at 30C therefore would be 6.92. The value of K.. at 0C
was given by Ackerman (1958) as 10-4.g5, which
means that neutral pH at that temperature is 7.48. The
strong effect of temperature on hydrogen-ion behavior
has considerable geochemical significance and must be
taken into account in measurements of pH and in calculations using pH data.
The hydrogen-ion content of a natural water computed in moles per liter (milligrams per liter for H is
nearly the same as millimoles per liter) is usually in the
trace constituent range. At pH 7, only 1~10.~ moles
per liter of hydrogen ion is present, for example. The
major constituents of most waters are in the concentration
range of 10. moles per liter and up. Thus the hydrogenion content does not begin to approach the status of a
major component of the solution until the pH goes below
4.0. A pH of less than 0 or greater than 14 can be attained
in concentrated acid or base solutions.
The hydrogen-ion activity in an aqueous solution is
controlled by interrelated chemical reactions that produce
or consume hydrogen ions. The dissociation equilibrium
for water is always applicable to any aqueous solution,
but many other equilibria and many nonequilibrium
reactions that occur in natural water among solute, solid
and gaseous, or other liquid species also involve hydrogen
ions. The pH of a natural water is a useful index of the
status of equilibrium reactions in which the water participates.
The reaction ofdissolved carbon dioxide with water,
which is one of the most important in establishing pH in
natural-water systems, is represented by the three steps
of Properties

and Constituents

Reported

in Water

Analyses

61

Adding these together gives

C02(g)+H20(1)=H2C03(aq),
H,C03(aq)=H+HC03-,
and

If a particular reaction or set of reactions involving


water, hydrogen ions, other solutes, and solids or gases
has attained equilibrium, the pH that should be attained
canbecomputedby the masslaw and a setof simultaneous
equations.Generalproceduresfor making such computations have already been outlined.
Under nonequilibrium conditions, the pH of the
natural water can reach a steady value effectively controlled by a single dominant chemical processor a set of
interrelated reactions. The controlling speciesare generally those present in the system or available to it in the
largest quantity or those whose reaction rates are fastest.
The control of pH by chemical equilibria can be
illustrated by a simplified example. When pure water is
in contact with a constant supply of gascontaining COZ
such asthe atmosphere,carbon dioxide will be dissolved
up to a specific solubility limit dependingon temperature
and pressure.At 25C and 1 atmosphere of pressure,the
chemical equilibria involved are those already given for
the solution and dissociation of dissolved carbon dioxide
and of water. Mass-law equations for these equilibria at
25C and 1 atmosphere are

HC03-=H++CO:-.
Both the secondand third stepsproduce H and influence
the pH of the solution. Other reactionsinvolving dissociations of acidic solutes include
HzP04-=HPO:-+H+,

H$(aq)=HS+H+,
and
HSOd-=SO:fH+.
Many of the reactions between water and solid
speciesconsume H; for example,

and
ZNaAlSi30B(c)+2Ht+9H20(1)
=AlzSiz05(OH)4(c)+4H,SiO~(aq)+2Na.

FLCO31
=q) =lo-1.43,

P
co,

The secondof theseinvolves alteration of water, and the


first could also be written as

WC%1

W+lzK

WzC031

=1o.6 35

CaC03(c)+HzO(l)=Ca2+HC03-+OH-.

[Hl [CO&
Theseare commonly called hydrolysis reactions, and all
such reactions influence or are influenced by pH.
Certain reactions in which precipitates are formed,
such as

[HCO;]

Also, the constraint of electrical neutrality requires that


there must be a cation-anion balance in the solution,
represented by Cn+=Cnco; +2Ccoi- +COH-. The
quantity PCC+is the partial pressureof carbon dioxide in
the gasphase-that is, the volume percent of CO2 multiplied by the total pressure,in atmospheres,and divided
by 100. The bracketed quantities are activities of solute
species, in moles per liter, and C terms represent ion
concentrations in the same units,.Becausethis will be a
very dilute solution, ion activities and actual concentrations are virtually the same.
There are five equations governing this system and
a total of six variables-the Pco, and five solute activities.

This reaction requiressomething to take up the electrons


it produces-that is, something must be reduced as the
sulfur is oxidized. The usual oxidizing agent probably is
atmospheric oxygen. The rest of the complete equation,
then, would be

and

Interpretation

of the

Chemical

Characteristics

[H] [ OH-]=~w=:l@4~00~

FeSz(c)+8H,0=Fe2++2S0:-+16H++14e

Study

and

influence pH, and most oxidation reactionsare pH related,

62

q=1o-lo33

of

Natural

Water

Therefore, if one variable is specified, all the others will


be fixed. The average value of Pco2 for the atmosphere is
lo- 353 and is nearly constant; therefore, pure water in
contact with air having the average CO2 content will
have a fixed pH and compatible concentrations of the
related ions. The pH calculated for these conditions is
5.65. At 0C the pH would be 5.60.
Although natural water can be expected to contain
other ions, rainwater, in the absence of atmospheric
pollution, might be expected to approach the conditions
specified in this example, and a pH in the vicinity of 5.6 is
frequently observed. The pH of ordinary laboratory distilled water often is near this value also.
If it is further specified that the system contains an
excess of solid calcite, CaCOa(c), the four mass-law
equations for equilibria still apply, along with a fifth
representing the solution of calcite:

[Cal WCGIEK
Wl

eq-

One more solute species is added to the ion balance


-equation, giving
2Cca2++cn+=cnco3+2cco;-+con-.
The system now has seven variables related by six equations, and, again, if one is specified all the others will be
fixed. For example, at equilibrium in this system, according to Garrels and Christ (1964, p. 83), if the gas
phase is ordinary air, the pH at 25C will be 8.4.
Garrels and Christ (1964, p. 83-91) gave calculations
for other conditions involving carbonates. It should be
evident from the examples how the pH enters into the
equilibrium calculations. Furthermore, it should be apparent that this way of studying natural-water composition has many interesting possibilities. It is unsafe, however, to assume that pH in natural water is always at a
value fixed by carbonate equilibria.
Buffered

Solutions

A solution is said to be buffered if its pH is not


greatly altered by the addition of moderate quantities of
acid or base. Buffering effects occur in systems in which
pH is controlled by reversible equilibria involving hydrogen and other ions, and the range of pH over which
buffering is effective depends on the nature of the other
solute species. Most natural waters are buffered to some
extent by reactions that involve dissolved carbon dioxide
species. The most effective buffering action by these
species is within the pH range of most natural waters.
Some insight into the mechanisms can be obtained
by briefly reviewing the carbon dioxide-water equilibria
Signiticance

just discussed. If the principal solutes are HzC03 and


HCOs- the first dissociation of carbonic acid,

will be the dominant control over pH. In mass-law form


at 25C and 1 atmosphere,
[HC03-]

=-

W&O,1

1O-635

Wl -

If strong acid is added to this system, the reaction will be


driven to the left, but a considerable number of hydrogen
ions may be used up before the ratio changes enough to
alter the pH noticeably. This buffering effect is strong in
solutions that contain considerable amounts of two dissolved CO2 species. In this example, if the total activity
of the two species is 2x10- moles/L, when their ratio is
unity, each will be present in an activity of lx 10m2
moles/L (610 mg/L HC0:3), and the pH will be 6.35. To
produce enough change in the ratio to cause a noticeable
decrease in pH (about 0.01 pH unit), one would have to
add at least 10m4moles/L of hydrogen ions to convert
HC03- ions to H2C03 species. The free hydrogen ions in
the system, as indicated by pH, would remain less than
10. moles/L. Thus the quantity of reacted H would be
some 100 times as great as the activity of free H.
Other equilibria within the system also are affected
when H is added or removed, and redistribution of the
H among all these other reactions adds to buffering
effects. The observed pH of a natural water is, in this
sense, a convenient indicator of the status of these equilibria, but H ions themselves normally are minor constituents of the solution. The pH of some waters is held at
levels considerably below the buffering range of dissolved
carbon dioxide species by redox reactions, such as the
oxidation of pyrite or sulfur in other forms.
Although the concept of control of pH by buffering
is related to the capacity of a system to participate in a
chemical reaction, the property of buffer intensity is
also a useful concept. For any given solution, this property
is defined as the reciprocal of the slope of the titration
curve at any given point in the titration (Stumm and
Morgan, 1981, p. 159).
The concept of buffering action can be extended to
heterogeneous systems, in which interactions between
solutes and solids may effectively maintain relatively
constant concentrations of H and other solute species
(Stumm and Morgan, 1981, p. 558).

Range

of pH

of Natural

Water

As stated previously, the pH of pure water at 25C


is 7.00. Most ground waters found in the United States
of Properties

and Constituents

Reported

in Water

Analyses

63

The pH of a water sample can also be affected by


oxidation of dissolved ferrous iron. The data in the
following table represent successivedeterminations of
ferrous iron and pH on a sample of water collected from
the overflow of the Senecaanthracite mine at Duryea,
Pa.,on July 16, 1963(unpub. analysesby U.S. Geological
Survey):

have pH valuesranging from about 6.0 to about 8.5, but


water having lower pH is not uncommon in thermal
springs. Water having a pH much greater than 9.0 is
unusual but by no means unknown. Values as high as
11.6 (Feth and others, 1961) and 12.0 (Barnes and
ONeil, 1969) have been observed in springs. The high
pH values of waters studied by Barnes and ONeil were
ascribed to reactions of meteoric water with ultramafic
rock in which serpentinitewas produced. Thesereactions
evidently consumeH more rapidly than it can be supplied
in this systemby any influx of carbon dioxide species.In
contrast, some thermal springs (for example, analysis 3,
table 18) may yield water whose pH is below 2.0.
River water in areas not influenced by pollution
generally has a pH in the range 6.5 to 8.5. Where
photosynthesisby aquatic organisms takes up dissolved
carbon dioxide during daylight and the organismsrelease
CO2 by respiration at night, pH fluctuation may occur
and the maximum pH value may sometimes reach as
high as 9.0. Livingstone (1963, p. G9) gave an example
of diurnal pH fluctuations in what was evident1y a poor1y
buffered lake water in which the maximum pH exceeded
12.
In carbonate-dominated systemsthe arrays of equilibrium equations cited earlier may provide a basis for
calculatingthe relative intensity of carbon dioxide sources.
This calculation requiresassumingthat measuredalkalinity and pH values can be used to calculate a partial
pressureof carbon dioxide in an initial source, such as
the air present in soil.
Table 11 gives three analysesof waters whose pH is
controlled by equilibria involving carbon dioxide species
and solid calcium carbonate. Analysis 1 is for a spring
whose water was rather highly charged with carbon
dioxide. The pH measured in the field at the time of
sampling was 7.54. The water precipitated calcium carbonate in the sample bottle, and a secondanalysis of the
sample made 6 months after the first shows a major
changein both calcium and bicarbonate concentrations,
accompanied by an increase in pH of almost a full unit.
Analysis 3 is for a spring issuing from limestone near the
mouth of the Little Colorado River in Arizona. This
water deposits travertine in the riverbed, and it is likely
that a pH value considerably higher than the 6.5 reported
in analysis 3 would be observed in the water where
precipitation was nearing completion. The pH of 9.4
reported for analysis 1, table 18, may be associatedwith
hydrolysis reactions involving silicates in a system containing little carbon dioxide. The pH of the sodium
carbonate brine from Wyoming (analysis 2, table 18)
probably was near 11, but it was not reported in the
published analysis.The low pH value (1.9) for analysis 3,
table 18, results from the oxidation of sulfur species.
Fumaroles in the vicinity yield sulfur dioxide.
64

Study

and

Interpretation

of the

Chemical

Characteristics

FP2

(w/L)
Dnw ( I963)

PH

July 16 _.........
July 30
..____
Aug. 20 _._.......
Sept. 24
Oct. 29 __........

0
14
35
70
105

135
87
41
2.2

4.98
3.98
3.05
2.81
2.69

The sample contained a high concentration of iron, and


its pH decreasedby more than 2 units as the ferrous iron
was oxidized and precipitated as ferric hydroxide.
Measurement

and

Interpretation

of pH

Electrochemical measurementof hydrogen-ion activity and the logarithmic pH scale for expressing the
results originated in the first decadeof the 20th century.
During the 1920s and 1930s, the glass electrode was
developed into a reliable and convenient sensingdevice
(Dole, 1941) and the determination of pH was widely
practicedin industrial and water treatment processcontrol
laboratories. Baylis (1927), for example, described the
useof pH data in treating water to decreaseits corrosivenesstoward iron pipe. A paper by Atkins (1930) pointed
out some possible applications in geochemistry, and
measured pH values appear in many water analyses
published during the 1930s.
A compilation of analyses (of water from the Rio
Grande at various points in Colorado, New Mexico and
Texas(Scofield, 1938)indicatesthat the U.S. Department
of Agriculture laboratory at Riverside, Calif., beganpublishing pH values as part of their water analysesin late
1933. U.S. Geological Survey laboratories did not begin
to publish such data for river water until about a decade
later. All these early measurements were made in the
laboratory, and instrumental field measurementsdid not
become common practice until the 1950s.
The design and construction of pH meters and
electrodeswere greatly improved during the 1950s and
1960s,and modern instrumentsare capableof measuring
pH either in the laboratory or in the field with an experimental reproducibility of t-O.01or 0.02 unit.This degree
of precision requires careful work and special attention
to electrode maintenance,buffer solutions, and temperature corrections. Barnes(1964) describedproceduresfor
precise field measurements.

of

Natural

Water

A more primitive method formerly used for measuring pH in the field employed colored pH indicators.
This technique is capable of measuring pH with an
experimental reproductibility of only +O. 10 unit under
the most favorable conditions and is not suitable for
water that is naturally colored or poorly buffered. For
the indicator compound to develop its color, it must
participate in a reaction using or generating H, and this,
in a nearly unbuffered solution, can overwhelm the
buffering system and give a false pH value.
The measured pH is a very important piece of
information in many types of geochemical equilibrium
or solubility calculations, and values that are inaccurate
or unrepresentative are a substantial source of error in
Table

11. Analyses

such work. Measurements of pH that date back to the


early 1950s or before may have rather large experimental
uncertainties owing to equipment or methods. In addition,
another serious type of uncertainty affects many pH
values that are reported in older water analysis literature.
Many of these values were determined in the laboratory
after the samples had been stored for a considerable
period of time. Field determinations of pH did not become
common practice in U.S. Geological Survey investigations
until the late 1960s. Obviously, field determinations are
much more likely to represent conditions in the water
that was sampled than are laboratory pH determinations.
Effects of temperature and mixing of water from different
aquifer sections in pumping wells may make interpreta-

of waters

with

unstable

pH

[Analyses by U.S. Geological Survey. Date under sample number ISdate of collection Sources of data, 1, 2, U.S
Geological Survey, unpubhshed data; 3, Cooley (1976, p. FIO)]

Constituent

Jan. 5, 1967

Jan. 5, 1967

June 14,195O

mg/L

meq/L

mg/L

meq/L

mg/L

19
22
22
....................
Silica (SiOn) ..........
.Ol
....................
.oo
....................
Iron (Fe) ................
.02
Manganese (Mn) ......
.Ol .................................................................................
Zinc (Zn) ..............
.02 ......................... .......................................................
264
4.64
22
1.10
Calcium (Ca) ........
93
79
4.77
54
4.44
Magnesium (Mg). ...
58
.65
513
.65
15
15
Sodium (Na) ..........
.03
23
.03
1.1
1.2
Potassium (K) ........
.Ol
.50
...........................................................
Strontium (Sr) ______
0
Carbonate (COa) ........................................................................................
Bicarbonate
964
9.02
297
4.81
556
(HCOa) ..................
147
1.04
55
1.15
50
Sulfate (Sod) ........
.48
815
.45
17
Chloride (Cl) ........
16
.2
.02 ..........................................
Fluoride (F) ..........
.3
3.2
.I
.oo ..........................................
Nitrate (N03) ........
Boron (B) ..............
.02 .................................................................................
Dissolved solids:
.... 2,340
..
332
Calculated ..........
530
Residue at
180C .......................................................................................................................
Hardness as
984
470
....................
217
....................
CaC0.3 ..................
194
....................
34
....................
Noncarbonate ____
19
3,940
Specific conduct...................................................................................
ante (micromhos
at 25C).
6.5
....................
8.40
....................
pH ____
......................
7.54

meq/L

13.17
6.50
22.31
.59

15.80
3.06
22.99
.Ol

.05

1. Spnng on headwaters of BlackbIrd Creek, SE1/4 sec.6, T. 6 S., R. 5 E., San Mateo County, Cahf Temperature,
7% pH measuredat ume of colleclion. Calcium determined on acldlfied sample.
2 Reanalysisof unacidified fraction of sample 1 after 6 months of storage in laboratory.
3. Spring discharging mto LIttIe Colorado River, I3 mi from mouth, Coconino County, Ariz One of a group known
as Blue Springs Water-bearing formanon Redwall Limestone. Water conlalns COn and deposus travertme
Discharge of group, about 200 cfs, temperature, 20 6OC pH measured in laboratory.

Significance

of Properties

and Constituents

Reported

in Water

Analyses

65

tion of the most carefully determined pH values a very


uncertain and difficult task.
As implied in the examples in this section, H ions
are produced within natural-water systems by various
types of chemical reactions and are largely consumed by
participating in subsequent chemical reactions in the
system. Graphical summaries of solute and solid phase
chemistry commonly use pH as one of the plotted variables, and some examples will be given in later sections
of this book.
Specific

Electrical

Physical

Pure liquid water has a very low electrical conductance: a few hundredths of a micromho per centimeter at
25C. This value has only theoretical significance, because
water this pure is very difficult to produce. The presence
of charged ionic species in solution makes the solution
conductive. As ion concentrations increase, conductance
of the solution increases; therefore, the conductance measurement provides an indication of ion concentration.
The relationship between ionic concentration and
specific conductance is fairly simple and direct in dilute
solutions of single salts. Figure 8 is a graph of the specific
conductance at 25C of potassium chloride solutions
with concentrations up to 0.01 molar (746 mg/L). The
relationship over this range of concentration is a straight
line. As the concentration is increased, however, the
slope decreases slightly; for a concentration of 7,460
mg/L, the conductance is 12,880 pmho/cm rather than
near 14,000, as an extrapolation of the slope in figure 8
would predict. This general behavior is typical of all salts,
but the slope of the straight part of the curve and the
degree to which it flattens with increasing concentration
are different for different salts.
Figure 9 shows the change in conductance of a
solution containing 746 mg/L of potassium chloride
between 0 and 35C. Over this temperature range the
conductance of the solution more than doubles. This
demonstrates the need for referring specific-conductance
measurements to a definite temperature. The response of
the conductance value to temperature change is somewhat
different for different salts and different concentrations,
but in dilute solutions for most ions an increase of 1C

Conductance

Electrical conductance, or conductiLity, is the ability


of a substance to conduct an electric current. Specific
electrical conductance is the conductance of a body of
unit length and unit cross section at a specified temperature. This term is synonymous with volume conductivity and is the reciprocal of volume resistivity (Weast,
1968, p. F-71). The American Society for Testing and
Materials (1964, p. 383) defined electrical conductivity
of water as the reciprocal of the resistance in ohms
measured between opposite faces of a centimeter cube of
an aqueous solution at a specified temperature. This
definition further specifies that units for reporting conductivity shall be micromhos per centimeter at tC.
Because the definition already specifies the dimensions
of the cube to which the measurement applies, the added
precaution of including the length in the unit may not be
essential and is often omitted in practice. Geophysical
measurements of resistivity, however, commonly are
expressed in ohm-meters, referring to a cube 1 m on a
side, so it may be a good idea to emphasize that conductantes of water refer to a centimeter cube. The standard
temperature for laboratory measurements is 25C, but
because other standard temperatures were used in the
past it is important that the temperature of measurement
be specified.
Units

for

Reporting

Study

and

Conductance

Because conductance is the reciprocal of resistance,


the units in which specific conductance is reported are
reciprocal ohms, or mhos. Natural waters have specific
conductances much less than I mho, and to avoid inconvenient decimals, data are reported in micromhos-that
is, the observed value of mhos is multiplied by 10.
Before October 1, 1947, the specific conductance values
reported by the U.S. Geological Survey were mhosx IO.
To convert these older values to micromhos, they should
be multiplied by 10.
Under the International System of Units (SI) it has
been proposed that the unit of conductivity by renamed
the siemens. The microsiemens (pS) is numerically the
same as the micromho. The change in terminology has
not yet been fully adopted in the U.S. literature.
66

Basis of Conductance

Interpretation

of the

Chemical

Characteristics

Figure

8. Specific

tions.
of

Natural

Water

conductance

of potassium

chloride

solu-

increases conductance by about 2 percent over the range


of temperature likely to be applicable to laboratory
conditions.
To conduct a current, solute ions actually must
move through the solution to transfer charges, and the
effectiveness of a particular ion in this process depends
on its charge, its size, the way it interacts with the solvent.
and other factors. The property called ionic mobility
represents the velocity of an ion in a potential gradient of
I V/cm. (It should be noted that this is not the same
property as the geochemical mobility of an element, a
concept relating to the ease of transport of elements in
geochemical cycles). Ionic mobility is decreased by increasing concentration owing to interferences and interactions among the ions. Limiting values attained in solutions of very low concentration for the major ions in
natural water are in the vicinity of 6x 10e4cm/s at 25C
(Glasstone and Lewis, 1960, p. 445).
It is apparent that even in rather simple solutions
the relationships that affect conductance may be complicated. More complete discussions of theory and application ofconductance such as those in textbooks of physical
chemistry (Glasstone and Lewis, 1960, p. 415-451) further emphasize this fact. Natural waters are not simple
solutions. They contain a variety of both ionic and undissociated species, and the amounts and proportions of
each may range widely. When applied to natural water,
therefore, the conductance determination cannot be expected to be simply related to ion concentrations or to
dissolved solids. Some water sources, however, can display well-defined relationships.
Figure IO is a plot of the dissolved-solids concentration in composite samples of water from the Gila River
at Bylas, Ariz., for a year (U.S. Geological Survey, 1947)
against the specific conductance of the samples. A reason-

ably well defined relationship is indicated for the range;


thus, for any given conductance value a dissolved-solids
value can be estimated with an uncertainty of only about
f 100 mg/L, using the curve, which was fitted by eye to
the points in figure 10. The whole set of data fit a straight
line regression closely (r=.98) with a value for A in the
formula

KA=S
of 0.59. (K in this formula is specific conductance in
pmhos/cm and Sis dissolved solids in mg/L.) However,
the data points show that the slope of the regression is
steeper for the lower conductance values. This is in
accord with the earlier discussion of the relationship of
conductivity to dissolved-solids concentration, which
indicated that A would not be constant over a wide
concentration range in solutions of a single anion and
cation species. This kind of formula is often used, however,
in calculating approximate dissolved-solids values from
conductance determinations. For the analyses of natural
waters given in this report, the range of A is about 0.54 to
0.96, which represents nearly the full range to be expected.
A is mostly between 0.55 and 0.75, the higher values
generally being associated with water high in sulfate
concentration.
Figure 11 shows the relationship of specific conductance to hardness and ion concentrations for the same set
of chemical analyses used in preparing figure IO. Rather
well defined relationships again exist for chloride and
sulfate, and almost as good a relationship for hardness
(calcium+magnesium) is indicated. The bicarbonate concentration of these solutions (not shown in fig. 11) is less
closely related to conductance, however. Lines drawn by

Figure

Specific conductance of O.Ol-molar solution of


potassium chloride at various temperatures.
Figure

9.

Significance

10.

composites
tober

of Properties

1,1943,

500

'

1000
DISSOLVED

Dissolved

solids

of daily

samples,

to September

and Constituents

'

'

'

'

3000
1500
2000
2500
SOLIDS, IN MILLIGRAMS PER LlTER

and
Cila

specific
River

conductance
at Bylas,

Ariz.,

3500

of
Oc-

30,1944.

Reported

in Water

Analyses

67

such uses of conductance data (can be made depends, of


couise, on ascertaining relationships like those in figures
10 and 11; however, this cannot be done for all streams
or all dissolved species of possible interest.
Conductance determinations are also useful in area1
extrapolation of ground-water analyses in areas where
comprehensive analyses are available for part of the
sampled points. Conductivity probes are of value in
exploring wells to determine differences in water quality
with depth.

eye through the points for chloride and sulfate show


slight curvature, but the departure from linearity is insignificant.
The data used in figures 10 and 11 were obtained
from a river that has a rather saline base flow maintained
by irrigation drainage and ground;water inflows. The
chemical characteristics of the base flow are rather constant and are subject to dilution by runoff. It seems
evident that a record of conductivity at this station could
be used to compute the other chemical characteristics of
the water with a good level of accuracy for major ions,
except at high flow when the relationships would not be
as well defined.
When a satisfactory set of relationships between
conductance and ion concentrations can be developed
from a few years of intensive sampling, a record of
major-ion concentrations, suitable for most uses, could
be obtained by measuring conductivity continuously
and feeding the output into a computer set up to convert
the data to concentrations averaged for whatever periods
might be desired. The sampling and analysis could be
directed toward determination of constituents that did
not correlate with conductance. The extent to which

Range of Conductance

Values

The specific conductance of the purest water that


can be made would approach 0.05 pmho/cm, but ordinary single-distilled water or water passed through a
deionizing exchange unit normally has a conductance of
at least 1.O pmho/cm. Carbon dioxide from the air in the
laboratory dissolves in distilled water that is open to the
air, and the resulting bicarbonate and hydrogen ions
impart most of the observed conductivity.
Rainwater has ample opportunity before touching
the Earth to dissolve gases from the air and also may

1600

EXPLANATION
C&ode

1000
SPECIFIC

2000
CONDUCTANCE,

IN

3000
MICROMHOS

PER

4000
CENTIMETER

AT

5000
25 DEGREES

CELSIUS

11. Relation of conductance to chlortde, hardness, and sulfate concentrations, Glla Roverat Bylas, Ariz., October 1,
1943, to September 30, 1944.
Figure

68

Study

and Interpretation

of the Chemical

Characteristics

of Natural

Wale8

dissolve particles of dust or other airborne material. As a


result, rain may have a conductance much higher than
distilled water, especially near the ocean or near sources
of atmosphericpollution. Feth and others(1964) reported
conductivities of melted snow in the Western United
Statesranging from about 2 to 42 pmho/cm. Whitehead
and Feth (1964) observedconductivity values of greater
than 100 in several rainstorms in Menlo Park, Calif.
The conductance of surface and ground waters has
a wide range,of course,and in some areasmay be aslow
as50 pmho/cm where precipitation is low in solutesand
rocks are resistantto attack. In other areas,conductances
of 50,000 pmho/cm or more may be reached;this is the
approximate conductance of seawater.Brine associated
with halite may contain as much as ten times the dissolved-solids concentration of seawater. Analysis 8 in
table 17representssuch a water; its specific conductance,
however, is 225,000 pmho/cm, about five times that of
seawater. At these high concentrations the correlation
between dissolved solids and conductivity is not well
defined.
Accuracy

and

Reproducibility

The equipment usually usedto obtain conductance


valuesof water, if carefully operated,may produce accuracy and precisionof from +2 to *5 percent.Instruments
that havea built-in capacity for compensatingfor temperature effects are available. The responseof specific conductance to temperature is different for different ions,
however, and thereforecompletely accuratetemperature
compensation applicable to all types of water is not
feasible. The best accuracy will be obtained by making
all measurementsat temperaturesnear 25C.
The accuracyof field measurementsof conductivity
madewith good portable equipment and proper attention
to temperature effects should be about the same as the
accuracy of laboratory measurementsif the temperature
is near 25C. Conductivity devicesthat havebeenpermanently installed in the field require periodic maintenance
to prevent electrode fouling and (or) other interferences
that may causeerroneous readings and loss of record.
A more comprehensivediscussion of conductance
theory and practice in water analysis hasbeenpublished
elsewhereby the writer (Hem, 1982).
Silica

The element silicon, asnoted earlier, is second only


to oxygen in abundancein the Earths crust. The chemical
bond between silicon and oxygen is very strong, and the
Si ion is the right size to fit closely within the space at
the center of a group of four closely packed oxygen ions.
Silicon thus located is said to be tetrahedrally coordinated
with respect to oxygen. The same structure occurs also
Significance

with hydroxide ions, which are nearly the same size as


the 02- ion. The SiOd- tetrahedron is a fundamental
building unit of most of the minerals making up igneous
and metamorphic rocks and is present in some form in
most other rocks and soils, as well as in natural water.
The term silica, meaning the oxide SiOz, is widely
usedin referring to silicon in natural water, but it should
be understood that the actual form is hydrated and is
more accurately representedas HASi or Si(OH)d.
The structure and composition of silicate minerals
cannot be consideredin detail here,but some knowledge
of the subject is useful in understanding the behavior of
silicon in natural water. There are six principal patterns
in which the SiO4 tetrahedra are joined to build up the
framework of silicate minerals. The kind of pattern that
occurs is a function of the relative abundanceof oxygen
in the rock compared with the abundance of silicon. In
systemsin which oxygen is abundant relative to silicon,
the predominant pattern is one in which adjacent tetrahedra are linked through chemical bonding of oxygen
with a divalent cation such as magnesium, for example,
in the magnesianolivine, forsterite (MgzSiO,). This pattern extendsin three dimensions,and silicates of this type
are called nesosilicates. A second structural pattern is
made up of pairs of tetrahedra sharing one oxygen ion
between them, the sorosilicates. Few natural minerals
have this structural pattern. A third pattern consists of
rings in which three or more tetrahedra each share two
oxygen ions. A six-member ring structure occurs in the
mineral beryl (Be3A1&a01s).The silicateshavingisolated
rings in their structure are called cyclosilicates. This
structure also is rather uncommon.
A structural patternin which eachtetrahedronshares
two oxygen ions with neighbors can produce a long
single chain as in the pyroxenes (&(SiOa)2). If two
adjacent chains are cross-linked by sharing some additional oxygen, the structure of the amphiboles is formed.
For thoseamphibolesin which no aluminum is substituted
for silicon and no univalent cations are present, the
composition may be expressedas R7(OH)2SisOa2.The
symbol R in these formulas representsvarious divalent
cations; most commonly theseare Ca, Mg2, and Fe2.
The chain silicates are called inosilicates.
The tetrahedra also may form planar structures in
which three oxygen ions of each tetrahedron are shared
with adjacenttetrahedra,all lying in a plane. The resulting
sheetstructure appearsin such mineral speciesas micas
and clays, for example kaolinite (AlaSiaOs(OH)4). The
sheet structures are called phyllosilicates. The tectosilicatesare made up of tetrahedra in which all oxygen ions
aresharedamongadjacenttetrahedrain threedimensions.
The structure of quartz (SiOz),, follows this pattern.
Feldsparsalso partially display it but contain somealuminum and other cations in place of silicon, asin potassium
feldspar (KAlSiaOs).
of Properties

and Constituents

Reported

in Water

Analyses

69

relatively easily compared with silicon-oxygen or aluminum-oxygenbonds.The ferroma.gnesianminerals, which


belong largely to thesetwo classesof silicate structures,
are lessresistant to weathering attack than are structures
like the tectosilicates in which silicon-oxygen bonding
predominates to a greater degree.
Crystalline SiO2 as quartz is a major constituent of
many igneous rocks and also constitutes the bulk of the
grainsin most sandstones.Of the common rock minerals,
quartz is one of the more resistantto attack by water. The
cryptocrystalline and amorphous forms of silica such as
chert and opal are more soluble. It seems probable,
however, that most of the dissolved silica observed in
natural water results originally from the chemical breakdown of silicate minerals in processesof weathering.
Theseprocessesare irreversible, and the silica retained in

The structural features of silicate minerals are described in greater detail by Hiickel (1950, p. 740-755)
and in many more recent texts on inorganic chemistry
and mineralogy. The structural pattern is rather closely
related to the stability of the various mineral species
when they are attacked by water. The silicon-oxygen
bond is stronger than the metal-oxygen bonds that occur
in the silicate minerals. Thus, the resistanceto chemical
attack, which involves breaking the bonds holding the
structure together, is greatest in mineral structures in
which a larger proportion of the bonds are between
silicon and oxygen.
The nesosilicatesand inosilicatesrepresentstructures
in which a relatively high proportion of the bonding is
the linking of divalent metal cations to oxygen. These
bonds representzonesof weaknessthat can be disrupted

Table 12. Analyses


[Analyses by U.S. Geological Survey. Date below sample number 1s date of collection.
(1954b. p. 286); 8, U.S GeologIcal Survey Water-Supply Paper 1102 (p. 400)]

1
Constituent

Nov.

23,

mg/L

99

Silica (SiOn) ............................


Aluminum

(Al)

Sources of data: 1, Scott and Barker (1962, p. 39); 2, 4, 5, and 9, U.S.

1953

Aug.

meq/L

....................

1, 1947

mg/L

Oct.

meq/L

103

....................

meq/L

363
.2
.06
.o
.8

............................................................................................................

..................................
.04 ....................
.o
....................
Manganese (Mn) ..........................................................................................................
2.4
,120
Calcium (Ca) ............................
6.5
,324
Magnesium
(Mg) ......................
1.4
,115
1.1
.090
.O
4.348
Sodium (Na) ............................
100
352
1.729
40
Potassium (K) ..........................
2.9
.074
24
i
Carbonate (CO,) ......................
24
.800
0
0
(7
111
Bicarbonate
(HCOa)
................
1.819
17
1.262
Sulfate (Sod) ............................
30
,625
23
15
,312
Chloride (Cl) ............................
10
,282
405
17
,479
Fluoride (F) ..............................
22
1.158
1.6
.084
25
Nitrate (N03) ..........................
.5
,008
.4
,006
1.8
Boron (B) _____........_...___............
.I9
.. . . .. . .. ..___...
.o
. . ..________._... 4.4

June 13,1952
mg/L

49
..........................
....................
.Ol
..............
.oo
.04
32
12
.oo
15.30
30
.61
5.2
3.86
0
220
.48
11
7.9
11.42
1.32
.2
.03
2.9
.._.________..... .08

meq/L

1.597
.987
1.305
,133
.O
3.605
,299
.223
.Oll
,047

solids:

Calculated
Residue

16, 1957

mg/L

Iron (Fe)

Dissolved

of waters

. . . ..__.__________........
on evaporation

348

Hardness
asCaC03_________.........
12
Noncarbonate ____....._.___________
0
Specific conductance
449
(micromhos

.._________
_...

222

. .. . .. .. .. . .. . ..

1,310

.______._____...........................................................................................................................

_________._.._______
20
0
167

_.._...____________.2
0
__.......___________
1,790

225
257
129
0
.. . .._____________
358

at 25OC).

9.2
. ...____________...6.7
._.___________......9.6
.__________......... 7.8
Color __............._____...............................................................................................................................................
10

pH

_.....___________.
. . . . . . _____________._.._

Includes some silicate ion, probably bSi0;.


calculated at 0.389 meq/L.
Original analysis reports carbonate and hydroxide alkaltnity of 3.86 meq/L

total, probably

mostly attrtbutable

to sdicate.

I. Flowtng well 7S 6E-9ba2, Owyhee County, Idaho. Depth, 960 ft; temperature, 50.0C.
2. Sprtng on the Rio San Antonio, SW1/4 sec. 7, T. 20 N., R. 4 E. (unsurveyed), Sandoval County, N. Mex. Flow, about 25 gpm; temperature, 38C Water from
rhyolite.
3. Spring, 650 ft south of Three Sisters Springs in Upper Geyser Basin, Yellowstone Nattonal Park, Wyo. Temperature, 94V. Also reported: 1.5 mg/L As, 5.2
mg/L Li, 1.5 mg/L Br, 0.3 mg/L I, 1.3 mg/L POi3, and 2.6 mg/L H2S.

70

Study

and

interpretation

of the

Chemical

Characteristics

of

Natural

Water

solution is probably controlled either by kinetic factors


in the dissolution process,adsorption on mineral or other
surfaces, or by precipitation of a secondary mineral
speciessuch as one of the less organized forms of silica.
This could be a clay mineral precursor that includes
other cationsaswell asthe silicon. The direct precipitation
of quartz is unlikely to control solubility of silica in most
natural waters at Earth surface temperatures.
Studiesrelating to weatheringmechanismsand silica
concentration controls made in recent years have suggestedand applied nonequilibrium models to good effect
(Helgeson, 1968, 1971;Busenbergand Clemency, 1976;
Paces,1978).

containing

high

proportions

Forms

of Dissolved

Silica

The convention of representing dissolved silica as


the oxide SiOa has been followed by all water analysts.
Some of the older literature implied that the material
was presentin colloidal form, but occasionally a charged
ionic species,SiOi-, was specified. In most waters, it is
fully evident that the dissolved silicon does not behave
like a chargedion. Nor doesit exhibit behavior typical of
a colloid in most waters.
The dissociationof silicic acid, which for this discussion may be assumedto have the formula Si(OH)4 or
H4Si04, begins with the reaction

of silica

Geological Survey, unpublished data; 3, White, Hem, and Waring (I 963, p. F40); 6, U.S. GeologIcal Survey Water-Supply

Apr.
mg/L

5
18, 1952
meq/L

38

6
Apr. 11 -19, 1944
mg/L

Oct.

meq/L

48

7
12, 1951

mg/L

Paper 1022 (p, 266); 7, Lohr and Love

8
May 6, 1947

meq/L

mg/L

.._.._____________71

9
22, 1952

Mar.

meq/L

mg/L

62

meq/L

29
.

.24
.oo
12
6.6
7.2
3.1
0
85
4.4
1.2
.3
.2

115
114
57
0
136
7.9
8

.03

.O
. .. . . .

.559
,543
,313
.079
.O
1.393
.092
.034
,016
,003

43
20
28
4.6
.O
254
8.3
28
.5
.l
0
306
190
0
463

2.15
1.64
1.22
.12
.O
4.15
.17
.I9
.03
.oo

.33
.oo

.Ol
.. .

1.098
.362

17
1.7

.... .. . . . ........
,848
,140

.870

7.4

.323

. . . .. . . . . . . . . . . . . . . . . . .

32
8.8

1.597
.I24

22
4.4

42

1.826

20

0
161
54
12

.O
2.639
1.124
,338
.037
,010

>

.I
.6

.O
104
29
0
.3
1.4

.O
1.704
.604
,000
,016
.023

0
.O
69
1.131
6.9
.144
1.1
.031
.l
.005
.._..____.___..._. .O

.
310
116
0
404

7.9
2

190
158
73
0
198

....................
....................
....................
....................
....................

98
98
49
0
130

....................
....................
....................
....................

8.1

4. Drdled well, NW114 sec. IO, T. 2 N., R. 32 E., Umatilla County, Oreg. Depth, 761 ft. Water from basalt of the Columbia River Group.
5. Flowmg well, SEl/4 set 16, T. I2 N., R. I7 E., Yaklma County, Wash. Depth, 1,078 ft; temperature, I7 2 OC. Water from basalt.
6 Eagle Creek at P-D pumping plant near Morencl, Greenlee County, Ariz Mean discharge for composite period, 9.9 cfs. Drainage basin ofabout 600-m? area in
which virtually all rocks exposed are extrusive volcanic rocks
7. Main pump statton, Albuquerque, N. Mex Public supply. Seven wells, 250 to 716 ft deep. Water from the Santa Fe FormatIon (valley till).
8. Mlddle Loup River at Dunning, Nebr. Discharge, 394 cfs Drainage from Nebraska sand hills, flow maintained by ground water.
9. Well at Valdese General Hospital, Rutherford College, Burke County, N C Depth, 400 ft; temperature, 15.0C. Water from mica schist.

Significance

of Properties

and Constituents

Reported

in Water

Analyses

71

The possible existence of fluosiiicate complex ions


in natural water was investigated by Roberson and
Barnes (1978), who found that substantiai proportions of
the silica was present as SiFi- in samples of condensed
gaseous material from fumaroies and drill holes near
Kiiauea Iki volcano, Hawaii. The amount of this species
in water associated with vuicanism elsewhere seemed
small, however, judging from applications of the same
model to data given by White and others (1963).

Values collected by Siiien and Marteii (1964, p. 145) for


the equilibrium constant for this reaction at 25C range
from 10. to lO-l. These values indicate that the first
dissociation step is half completed at a pH value somewhere between 9.41 and 9.91, and silicate ions (H3Si04)
might constitute no more than 10 percent of the total
dissolved silica species at a pH between 8.41 and 8.91.
Any silicate ions that might be present are converted to
siiicic acid in the alkalinity titration and appear in the
analysis as an equivalent amount of carbonate or bicarbonate. As a result, the ionic balance is maintained in the
analysis, and there is no obvious indication that the ionic
species are incorrectly identified. In any detailed interpretation of the chemistry of high-pH waters, the contribution
of H3Si04- to the alkalinity must be taken into account.
Analysis 1, table 12, is for a water that had a pH of
9.2. By using the dissociation constant for siiicic acid
determined by Greenberg and Price (1957) of 10. 71, it
can be calculated that about 0.389 meq/L of dissociated
silicate ions are present. The analysis reports 0.800 meq/L
of carbonate alkalinity, of which almost half actually
must have been caused by silicate.
The method generally used for determining silica
concentrations in water requires that the silica form a
colored moiybdate complex in solution. The response of
dissolved silica to this procedure is normally rapid and
apparently complete, which suggests that silica must be
present in units of very small size, probably approaching,
if not actually having, the dimensions of single molecules.
Some hot-spring waters are known, however, in which
the total silica is several hundred

milligrams

Solubility

Controls

The soiubiiity of quartz has been determined by


Morey and others (1962) as 6.0 mg/L at 25C and 26
mg/L at 84C (as SiO2). Fournier and Rowe (1962)
reported the soiubiiity of cristobaiite to be 27 mg/L at
25C and 94 mg/L at 84C. Morey and others (1964)
reported the soiubiiity of amorphous silica to be 115
mg/L at 25C, and Akabane and Kurosawa (1958)
reported the soiubiiity of this material at 100C to be 370
mg/L. Fournier and Rowe (1966) suggested that the
silica content of water from hot springs could be used to
calculate the temperature of reservoir rocks, assuming an
equilibrium with quartz at depth. This method has been
widely used in evaluating geothermal resources. The
soiubiiity of quartz over the temperature range 25C to
900C and at pressures up to 10,000 bars was later
determined by Fournier and Potter (1982), and the effects
of sodium chloride in solution on these relationships
were evaluated by Fournier and others (1982).
The first precipitate formed on cooling a silica-rich

per liter and

solution

to 25C is amorphous

silica, and this material

is

in which the silica tends to polymerize on standing at


25C to form colloidal particles. White and others (1956)
observed that the polymerized silica thus formed reacted
very slowly with compiexing reagents. Analysis 3, table
12, represents hot-spring water that precipitates silica on
cooling.
In reviewing the available literature on the dissoiution of silica, Krauskopf (1956) concluded that silica in
natural water was mostly in the form of monomolecular
siiicic acid, H4Si04(aq). This is equivalent to a silicon
ion tetrahedraiiy coordinated with four hydroxide ions,
and by analogy with mineral structures it seemed the
most likely form. Some investigators quoted by Siiien

not readily converted at this temperature to the better


crystallized forms of lower soiubiiity. Morey and others
(1962) reported one experiment in which quartz grains
were rotated in water for 386 days at 25C, during which
time the SiOn concentration in the water was 80 mg/L.
The concentration dropped suddenly to 6 mg/L and
held that value for about 5 months. It was concluded this
was the reversible soiubiiity for quartz.
Natural water normally has more dissolved silica
than the quartz equilibrium value but less than the amorphous silica value, which is the likely upper equilibrium
limit. The general tendency for silica concentrations in
natural water to fail within a rather narrow range suggests

and Martell

that some other type of solubility

(1964,

p. 145) suggest polynuclear

species

72

Study

and

Interpretation

of the

Chemical

Characteristics

control

may exist. Hem

and others (1973) suggested that an amorphous clay


mineral having the composition of haiioysite could be
produced by weathering igneous rock minerals and that
an equilibrium with such a species may control aluminum
and silica concentrations in some natural water. A somewhat similar mechanism was proposed by Paces (1978).
The dissolution of the sodium feldspar, aibite, can
be summarized

such as Si40d(OH):j,
and Iier (1955, p. 19) proposed
that in solution silica might be coordinated with six OHions to form ions of the same type as the fluosiiicate ion,
SiFE-. In the absence of more definite evidence of the
existence of more complicated species, the simplest form,
Si(OH),, is probably the best postulate. This may also be
considered equivalent to SiOg combined with two water
molecules.
of

Natural

Waten

=Al&OS(OH),(c)+4H,Si0,(aq)+2Na.
If the clay mineral produced dissolves reversibly,
Al&05(OH),+6H=2H,Si0,(aq)+2A13+H,0.
An upper limit for silica concentration

might be fixed by

H4Si04(aq)=SiOz(c) (amorph)+2Hz0.
These reactions are oversimplified. A higher Si:AI
ratio in the clay mineral, admixtures of other cations, and
many other complications can occur. Adsorption of
silica on other mineral surfaces can also be a factor, as
noted by Siever and Woodford (1973, 1979).
The kinetics of dissolution of igneous rock minerals
have been studied extensively in recent years. Among
papers on feldspar are those of Busenberg and Clemency
(1976) Petrovic (1976), and Holdren and Berner (1979).
Studies of inosilicate minerals were made by Lute and
others (1972), Siever and Woodford (1979) and other
groups. Rate-controlling mechanisms of various types
have been proposed to explain the experimental data.
Kinetics of dissolution and precipitation of quartz and
three other forms of solid SiOz from 0 to 300C were
determined by Rimstidt and Barnes (1980).
Occurrence

in Natural

Water

Analyses in table I2 were selected to show several


types of water that have rather high concentrations of
dissolved silica. Some of these show the increased solubility attained as water temperatures rise. The highest
concentration in natural water encountered by the writer
was reported by Feth and others (1961) for Aqua de
Ney, a mineralized cold spring near the town of Mount
Shasta, Calif. This water has a pH of 11.6 and more than
3,400 mg/L SiOZ, presumably mostly as dissociated
silicate.
Seawater near the surface is very low in silica (commonly less than 1 mg/L), apparently because marine
organisms (primarily diatoms) extract and use silica in
their shells and skeletons. Similar depletion effects are
noticeable in the surface layers of some lakes and reservoirs.
The range of concentrations of silica most commonly
observed in natural water is from I to about 30 mg/L.
Concentrations up to 100 mg/L, however, are not infrequent in ground water in some areas. Davis (1964)
quoted a median value of silica for surface waters of 14
mg/L and for ground water of I7 mg/L. The higher
concentrations found in ground waters are related to
rock type and to temperature of the water. A general
review of silica occurrence in natural water has been
Significance

published by Ginzburg and Kabanova (1960). Some


geologic factors related to silica concentrations in water
are considered elsewhere in this volume.
Inspection of analyses of surface waters of the world
assembled by Livingstone (1963) indicates that most
streams in the Northeastern United States carry water
containing less than 10 mg/L of silica, but water from
drainage basins in the West and from some in the South
seem generally to be higher.
The silica concentration of water samples collected
and stored in some types of glass bottles may increase if
stored for a long time owing to solution of the glass.
Bottles made of Pyrex and other resistant formulations of
glass probably are not a serious source of contamination
for most natural waters, however. Polyethylene and polypropylene bottles are superior to glass in this respect.
Aluminum

Although aluminum is the third most abundant


element in the Earths outer crust, it rarely occurs in
solution in natural water in concentrations greater than a
few tenths or hundredths of a milligram per liter. The
exceptions are mostly waters of very low pH. Because
aluminum is so abundant and so widely distributed, most
natural waters have ample opportunity to dissolve it. The
low concentrations common to water at near-neutral pH
must, therefore, be a result of the chemistry of the element.
These chemical properties are also indicated by aluminums short oceanic residence time (table 7).
Sources

of

Aluminum

in Water

Aluminum occurs in substantial amounts in many


silicate igneous rock minerals such as the feldspars, the
feldspathoids, the micas, and many amphiboles. The
aluminum ion is small enough to fit approximately into
fourfold coordination with oxygen and therefore can
substitute, in a sense, for silicon in tetrahedral structural
sites. It also is commonly six-coordinated, and occupies
octahedral crystal sites similar to those occupied by
magnesium and iron. Aluminum is trivalent, and substitutions of these kinds may require adding or removing
cations or protons to maintain a net charge balance in the
structure.
On weathering of igneous rocks, the aluminum is
mostly retained in new solid species, some of which may
be greatly enriched in aluminum. Nearly pure aluminum
hydroxide in the form of gibbsite is a fairly common
mineral, and less common hydroxides include nordstrandite and bayerite, whose composition and structure
are similar to that of gibbsite. In low-pH environments
aluminum may be precipitated as an aluminum hydroxysulfate. The most common of the sedimentary aluminumenriched minerals are clays. The clay minerals have a
layered structure in which aluminum octahedrally coof Properties

and Constituents

Reported

in Water

Analyses

73

ordinated with six oxide or hydroxide ions (gibbsite


structure)forms one type of layer, and silicon tetrahedrally
coordinated with oxygen forms a second type of layer.
Theselayersalternatein various ways,forming the various
clay structures.The layersare bound togetherby Si-O-Al
bonds. Diagrams of the structures of the common clay
mineralshavebeengiven by many investigators,including
Grim (1968, p. 51-125) and Robinson (1962, p. 31-58).
Clays are present in most natural-water environments
and are abundant in most soils and in hydrolyzate sedimentary rocks.
Species

in Solution

The cation A13 predominates in many solutions in


which pH is lessthan 4.0. The actual ion is probably an
octahedron of six water molecules with a aluminum ion
at the center (fig. 12). One of thesewater molecules may
become an OH- ion if the pH is raised slightly, and at pH
4.5-6.5 a processof polymerization occurs that results in
units of various sizeswith the structural pattern of gibbsite
(Hsu and Bates, 1964; Hem and Roberson, 1967). This
structure is characterizedby hexagonalrings of aluminum
ions bound together by sharing pairs of hydroxide ions.
Under some conditions the polymerization may stop
when the units are still small, but in studies described by
Smith and Hem (1972) the polymers continued growing
until they became crystalline gibbsite particles a few
hundredths to a few tenths of a micrometer in diameter.
Above neutral pH, the predominant dissolved form of
aluminum is the anion Al(O
In someof the published researchon aqueouschemistry of aluminum, certain specific aluminum hydroxide
polymeric ions are suggestedas the predominant forms
to be expected in solution in natural water. The work on
aluminum hydrolysis specieswas summarized by Baes
and Mesmer (1976,112-123). As noted above, there are
some conditions under which the polymeric ions may
stop growing. However, they should be viewed as precursors of a crystalline solid of well-defined structure
with which the form of the ions must be compatible. This
requirement is not met by some of the proposed species.
Further, if the polymeric ions increase in size with time
they are inappropriate for consideration as chemicalequilibrium species.
The polymerization of aluminum hydroxide species
proceedsin a different way in the presenceof dissolved
silica than when silica is absent.When sufficient silica is
present,the aluminum appearsto be precipitated rapidly
as rather poorly crystallized clay-mineral species(Hem
and others, 1973). In the presenceof an organic solute
capableof complexing aluminum, thesemixed silica and
aluminum solutions produced well-crystallized kaolinite
on aging at 25OC(Hem and Lind, 1974).
In the presence of fluoride, strong complexes of
aluminum and fluoride are formed. The ions AlF and
74

Study

and Interpretation

of the Chemical

Characteristics

AIF; appearto be most likely in natural water containing


from a few tenths of a milligram per liter to a few
milligrams per liter of fluoride (Hem, 1968b). Soluble
phosphate complexes of aluminum have been reported
(Sillen and Martell, 1964,p. 186) and the sulfate complex
AlSOi may predominate in acid solutions in which
much sulfate is present (Hem, 1968b).
Organic complexing of aluminum apparently occurs
in somenatural waters that are colored owing to solution
of humic material. Lind and Hem (1975) quoted analysesof 28 such solutions taken from published records for
U.S. streams,in which aluminum concentrations ranged
between 100 and 1,300 pg/L. One sample contained 38
mg/L and had a pH of 9.4, owing, however, to direct
pollution from bauxite mines and alumina processing
plants (Halbert and others, 1968). All these samples
were filtered through 0.45-pm porosity filters and some
may have contained colloidal forms of aluminum that
passedthrough the filters, as Kennedy and others (1974)
observed to occur in their studies of the effectivenes of
filtration techniques.
Solubility

Controls

Application of equilibrium models to aluminum


behavior requires special attention to identifying the
form of dissolvedspeciesand specialsampling techniques.
This may entail adding some analytical reagents and
performing part of the determination in the field (Barnes,
1975). In solutions that contain fluoride, sulfate, or other
complexing agents,the form of dissolved speciescan be
determined from the ionic equilibria (Hem, 1968b), and

Figure

12. Schematic

AI(H20)63i.
Water,
can
of Natural

(From
Advances

Chemical
Water

diagram
Hem,

in Chemistry

Society,

of hydrated

196810,

1968,

in

Series,
used

Trace

aluminum

Ion,

lnorganics

73, copyright

by permission.)

Ameri-

in

ifsilica concentrationsare very low, the solubility product


for aluminum hydroxide as bayerite gives a reasonable
basis for calculating solubility of aluminum in alkaline
solutions (Hem and Roberson, 1967). In solutions whose
pH is below 4.0, the solubility usually can be calculated
from the solubility product for gibbsite.Nordstrom (1982)
hassuggestedthat aluminum sulfate and hydroxysulfate
minerals may control the solubility of the element in
acidic waters that contain sulfate. Between the high and
low pH regions,aluminum solubility reachesa minimum.
The minimum solubility calculated by Roberson and
Hem (1969) is a little lessthan 10 r.lg/L, near a pH of 6.0.
May and others(1979) determined a minimum solubility
of 6.7 pg/L for synthetic gibbsite at pH 6.0 and about 27
pg/L for natural gibbsite at the same pH. These data
would be applicable at 2YC to a solution with a low
dissolved-solidsconcentration containing lessthan 0.10
mg/L of fluoride. Higher fluoride concentrationsincrease
the solubility of aluminum. Figure 13 is a graph of
solubility of aluminum as a function of pH in a system
that has no complexing species other than OH. The
polymeric forms of aluminum and hydroxide cannot be
represented in a solubility diagram, because they are
unstableand are converted to solid particulate matter on
aging(Hem and Roberson,1967).The dashedline labeled
microcrystalline gibbsite represents the equilibrium
activity of uncomplexed aluminum, and the other two
dashed lines represent the equilibrium activity of
Al(O
The solid line showsthe effect of including the
monomeric complex AIOH2 in the solubility calculation.
I

1
40

50

SO

70

I
SO

on

Figure

13. Equilibrium

formsofaluminum
activity

of Ai3+AIOH2

activities
hydroxide
(solid

of free
(dashed

aluminum
lines)

and

for three
calculated

line).
Significance

90

Although the influence of silica cannot be fully


evaluatedby equilibrium calculations, aluminum solubility generally is considerably decreased when silica is
presentowing to formation of clay-mineral species.The
equilibrium activity of aluminum in uncomplexed form
in the presenceof kaolinite or halloysite can be calculated
for various silica concentrations and pH values by using
equations given by Polzer and Hem (1965) and by Hem
and others (1973) and may be lessthan 1.Opg/L in some
natural waters.This concentration, however, is not equivalent to the total possible dissolved aluminum, which
could be much greaterif complexing anions were present.
Occurrence

of Aluminum

in Water

Commonly, aluminum determinations are not included in general-purposewater analyses,and the reliability of the limited amount of information that is available
is questionable. Occasional reported concentrations of
1.0 mg/L or more in water having near-neutral pH and
no unusual concentrations of complexing ions probably
represent particulate material. Whether this particulate
material is aluminum hydroxide or an aluminosilicate is
not presently known. The writers work (Hem and
Roberson, 1967) hasshown, however, that gibbsite crystals near 0.10 pm in diameter haveconsiderablephysical
and chemical stability. Particles of this size will pass
through most filter media and may needto be considered
in water-quality evaluations.Carefully detemined aluminum values for runoff from granitic terrane were only a
few hundredthsof a milligram per liter at most (Feth and
others, 1964). These determinations probably did not
include particulate material.
The addition of aluminum sulfate (alum) in watertreatment processesto flocculate suspended particles
may leave a residue of aluminum in the finished water,
probably as small, suspendedhydroxide particles. For a
study of quality of water supplies of major U.S. cities
(Durfor and Becker, 1964),spectrographicanalyseswere
madeof various minor constituentsincluding aluminum.
The aluminum concentrations commonly were higher in
finished water after alum had been usedasa flocculating
agent than they were in the original untreated water.
More recently, Miller and others (1984) observed a
similar effect in a study of water samplesfrom 186 water
utility systemsin the United States.
Water having a pH below 4.0 may contain several
hundredor evenseveralthousandmilligrams of aluminum
per liter. Such water occurs in some springs and in
drainagefrom mines. Table I3 contains analysesof some
watersthat are high in aluminum. Someof the commonly
used analytical proceduresare not sensitive to all polymeric aluminum species,and this further decreasesthe
usefulnessof analytical data when methodology is unknown.
of Properties

and Constituents

Reported

in Water

Analyses

75

low pH has a deleterious effect on fish and some other


forms of aquatic life (Driscoll and others, 1980).
The aluminum concentrations in analyses 2,3, and
9 in table I4 and I, 8, and 10 in table 16, which range
from 0.6 to 1.4 mg/L. probably can be ascribed to
polymeric colloidal material.

Analyses 1 and 4 in table 13 represent water of low


pH, and the thermal spring represented by analysis 3 had
a rather high fluoride concentration. The explanation for
the aluminum concentrations in the other two samples
may be related to the presence of particulates or polymers,
but neither analysis 2 nor 5 exhibits high concentrations.
Elevated aluminum concentrations have been observed in runoff and lake waters in areas affected by
precipitation having a low pH (acid rain). An occurrence in New England was described by Johnson and
others (1981). The dissolved aluminum in waters having
Table 13. Analyses

of waters

high

Iron

Although iron is the second most abundant metallic


element in the Earths outer crust, concentrations present
in dissolved

aluminum

or manganese

[Analyses by U.S Geolog~al Survey Date under sample number IS date of collectIon Sources of data I and 2. Scott and Barker (1962. p. 23. Iii).
Geologvzal Survey, unpubhshed data; 4, U.S. GeologIcal Survey Water-Supply Paper 1948 (p 38); 5, White. Hem, and Warmg (1963. p F26)]
1
Dec. 13, 1955

Constituent

Aluminum

............

98

........

28

(Al)

Iron (Fe)

.................

Manganese
Calcium

(Mn)

(Ca)

Magnesium
Sodium

...........

(Mg)

(Na)

.....

...........

Aug.

meq/L

9.6

.................

.35

................

.04

..................

.02

.....

..................

.31

................

21.16

58

2 89

194

15.96

13

I 07

416

18.10

23

I 00

Potassium

(K)

.........

(H)

.........................

Carbonate

(CO:j)

.28

I1
0

477

..................

mg/L

9.7

...............

31

3.5

..................

.2

.................

2 7

.I0
2.5

meq/L

.22

................
............
................

32

I .tjO

.oo

II

.oo

1.9

.I6

12

52

6.8

.30

4.2

II

20.75

..................

5
25, 1953

3.34

28
.07
40
1.02
..........................................................................................

10

......

.O

Mar.

meq/L

..........

67

4
8-lo,1963-

mg/L

92
.................

424

Hydrogen

Mar. l-6,

I
1.3

................

3
31, 1958
meq/L

mg/L

10
.88

.....

mg/L

meq/L

mg/L
Sihca (SiOy)

2
Dec. 3, 1955

3. U S

3.7

28

.09
I6

.oo

I 40

...................................

Bicarbonate
(HCO:r)

...................

Sulfate

(Sod)

............

(Cl)

............

(NO:{)

..........

Chloride

380
(F) ______________ 1.8

Fluoride
Nitrate

101

1.66

1,020

16.72

50.38

116

2.42

169

3.52

10.72

39

206
6.8

5.81
.36

18

.03

0
2,420

3.1

.09

.o

.05

.6

1.10
..................
.Ol

I21

0
I71

1.98

3.56

14

.03

5.0
.I

.:4
.Ol

I.0
.I

.03
.Ol

5.3

.09

.2

.oo

Orthophosphate
(Pod)

..... ..................

Boron

(B)

Dissolved
Residue
180C
CaC03

.I

.I1

..................

..................................

2.8

........................................................
.__

.....................................

.o

..................

... .........................................

...........

3,990

..

..................

4,190

...

314

.................

1,570

..

256

.................

137

...

260

.....................................................

..

125

.................

at
338

.....

1,560

I98

..

I68

as
......................

Noncarbonate
Specific

..................
.......................

solids:

Calculated

Hardness

.o

........ ........................

......

conductance

(micromhos

..

1,860
1,860

I15

.................

4,570

...

517

......

..................

2,430

78

I25
...

0
..

507

192

..............

at

25T).
pH

............................

4.0

..................

7.0

..................

6.7

..................

3.8

..................

6.9

I. Well, 7 ml northeast of Monttcello, Drew County, Ark. Depth, 22 ft. Water-bearmg formanon, shale, sand, and marl of the Jackson Group. Also contamed
radium (Ra), 1 7 pCl/L, and uramum (U), I7 &L.
2. Composite from two radial collector wells at Parkersburg, Kanawha County, W. Va Depth, 52 ft Water from sand and gravel Also contamed copper (Cu). 0 01
mg/L.

and nnc

(Zn),

0 01 mg/L

3. Wagon Wheel Gap hot spring, Mlneral County. Co10 Discharge, 20 gpm; temperature, 62.2C. Associated wnh wn of Ihe Wagon Wheel Gap fluorue mine
Also contamed 2.3 mg/L LI, 0 9 mg/L NH,, 0 3 mg/L Br, and 0 3 mg/L I
4 Klsklmmltas River at Leechburg (Vandergrift), Pa. Composite of mne dally samples Mean chscharge for period. 10,880 cfs.
5 Well. 167 ft deep, Balnmore Counly. Md. Water-bearlng formanon, Port Deposn granlnc gnelss Also contamed 0 01 mg/L copper (Cu).

76

Study

and Interpretation

of the Chemical

Characteristics

of Natural

Water

in water generally are small. The chemical behavior of


iron and its solubility in water depend strongly on the
oxidation intensity in the systemin which it occurs; pH is
a strong influence as well. The chemistry of iron in
aqueous systems has been studied extensively, and its
general outline can be readily discerned by applications
of equilibrium chemical principles.
Iron is an essential element in the metabolism of
animals and plants. If present in water in excessive
amounts,however,it forms red oxyhydroxide precipitates
that stain laundry and plumbing fixtures and, therefore,
is an objectionable impurity in domestic and industrial
water supplies. For this reason, iron determinations are
commonly included in chemical analyses of water. A
recommendedupperlimit for iron in public water supplies
is 0.3 mg/L (NAS-NAE, 1972).
Sources

of Iron

Igneous rock minerals whose iron content is relatively high include the pyroxenes,the amphiboles,biotite,
magnetite, and, especially, the nesosilicate olivine. The
latter is essentially a solid solution whose end members
are forsterite (Mg,Si04) and fayalite (FezSiOd). For the
most part, iron in these minerals is in the ferrous, Fe+,
oxidation state, but ferric (Fe) may also be present, as
in magnetite, FeaOq.
When these minerals are attacked by water, the
iron that may be released is generally reprecipitated
nearbyassedimentaryspecies.Under reducingconditions
when sulfur is available, the ferrous polysulfides such as
pyrite, marcasite,and the lessstablespeciesmackinawite
and griegite may occur. Where sulfur is less abundant,
siderite (FeCOs) may form. In oxidizing environments
the sedimentary specieswill be ferric oxides or oxyhydroxides such as hematite, Fez03, goethite, FeOOH, or
other minerals having these compositions. Freshly precipitated material may have poorly developed crystal
structure and is commonly designatedferric hydroxide,
Fe(OH)s.
Magnetite tends to resist attack by water and is
commonly present as a residue in resistate sediments.
Iron is a common constituent of sulfide ores of other
metals, and ferrous sulfide is generally associated with
coal seams.
Availability of iron for aqueoussolution is strikingly
affectedby environmental conditions, especially changes
in degree or intensity of oxidation or reduction. High
concentrations of dissolved ferrous iron can occur in
solution at the sites of either reduction of ferric oxyhydroxides or oxidation of ferrous sulfides. In the latter
processthe sulfur is attacked first and altered to sulfate,
releasing the ferrous iron. Iron is present in organic
wastesand in plant debris in soils, and the activities in the
biospheremay havea strong influence on the occurrence
of iron in water. Micro-organismsare commonly involved
Significance

in processesof oxidation and reduction of iron, and some


speciesmay use these reactions as energy sources.
Species of Iron in Natural

Water

The most common form of iron in solution in


ground water is the ferrous ion Fe. Like aluminum and
many other metal ions, this ion hasan octahedral hydration shell of six water molecules.Except where hydration
may have special significance, water of hydration shells
will not be shown in the formula of this or other ions that
follow. The monohydroxide complex FeOH can be
predominant abovea pH of about 9.5 (Baesand Mesmer,
1976, p. 100) and may be significant at somewhat lower
pHs. Above a pH of 11, the anion Fe(OHS)- or HFe02can exist in water in appreciableconcentrations, but such
a high pH is rarely attained in natural systems. An ion
pair FeSO.,(aq) reported in Sillen and Martell (1964, p.
240) could be important in solutions that have more than
a few hundred milligrams of sulfate per liter. Ferrous
complexes are formed by many organic molecules, and
someof the complexesmay beconsiderablymore resistant
to oxidation than free ferrous ions would be. Organic
compounds containing iron are particularly important in
life processes,suchasphotosynthesis,and in the functions
of hemoglobin in the blood of animals.
Ferric iron can occur in acid solutions as Fe3,
FeOH, and Fe(OH)i and in polymeric hydroxide
forms, the predominant form and concentration depending on the pH. Above a pH of 4.8, however,the total
activity of these species in equilibrium with ferric
hydroxide will belessthan 10pg/L. Dimeric or polymeric
ferric hydroxy cations (Fez(OH); , etc.) can become
important in solutions in which total dissolved ferric iron
exceeds1,000 mg/L. Such iron concentrations are rare
in natural water. However, a processof polymerization
somewhat like the one describedfor aluminum precedes
the formation ofsolid ferric oxyhydroxide, and macroions
or microcrystalline forms approaching the composition
Fe(OH)3 may often be present in natural water at very
low concentrations. The species Fe(OH)s(aq) that is
sometimes reported to be among the forms of dissolved
ferric iron may actually represent such polymers. The
stability of this material asreported by Lamb and Jacques
(1938) limits the equilibrium concentration to less than
1.0 /*g/L in the presenceof ferric hydroxide.
A AGP value for the speciesFeOd2-in which iron is
in the 6+ oxidation state was given by Garrels and Christ
(1964, p. 413). This form of iron is not likely to be
important in natural water becauseit could dominate
only in very strongly oxidizing systemsat high pH, and it
is not included in the pH-Eh diagrams in this book.
The only anionic ferric speciesfor which data are
available is Fe(OH)/. Langmuir (1969b) indicated that
his value for its AGf was a rough estimate. The value
used here was calculated from a stability constant estiof Properties

and Constituents

Reported

in Water

Analyses

77

mated by Baesand Mesmer (1976, p. 104). The effect of


Fe(OH)- on iron solubility is not significant, in any
event, unlessthe pH is 10 or greater.
Ferric iron forms inorganic solution complexes with
many anionsbesidesOH-. The chloride, fluoride, sulfate,
and phosphate complexes may be important in some
natural systems.Organic complexes containing iron are
presentin important amounts in some waters. Complexes
can occur either with ferrous iron (Theis and Singer,
1974) or ferric iron (Stumm and Morgan, 1970,p. 531).
The latter form may be associatedwith organic colloids
or humic-type material that gives some waters a yellow
or brown color. Such associationsmay be important in
trapping iron in estuaries by a process of coagulation
(Sholkovitz, 1976).
Ferric oxyhydroxide surfaces have a substantial
adsorption capacity which may affect the concentration
of minor constituents of water associated with such
material. Redox coprecipitation processes may occur
that can control solubilities of other metal ions under
some conditions (Hem, 1977a).
Solubility

Calculations:

pH-Eh

Diagrams

The chemical behavior of iron can be predicted


theoretically as a function of solution pH, oxidation (or
redox) potential, and activity of other ions. A convenient
summarizing technique that hasbeen usedextensively in
publications on iron chemistry is the pH-Eh (or pHredox potential) diagram. This is a two-dimensional
graph in which pH is plotted on the abscissaand redox
potential on the ordinate. From tables of chemical thermodynamic data applied to the Nernst equation and
pertinent mass-law expressions, an array of equations
can be developed.If enough of the variablesare specified,
the equations can be solved simultaneously, leading to
relationships expressedas points or lines on the pH-Eh
grid.
The pH-Eh diagram wes extensively developed by
Marcel Pourbaix in Belgium in the years prior to World
War II. The diagrams began to be used in geochemistry
after a compilation (Pourbaix, 1949) and applications by
R.M. Garrels(1960) in the United Statesmade Pourbaixs
work more widely known. Details of preparation of the
diagrams are given in various publications (Hem and
Cropper, 1959; Garrels and Christ, 1964; Hem, 1965).
Application of these diagrams to iron systems has been
used as a textbook example, and a general agreement
between theoretical calculations and field behavior of
iron hasbeendemonstratedby Barnesand Back (1964b)
and many other investigators. Experience,therefore, suggeststhat reactions representedin the theoretical model
for iron commonly approach chemical equilibrium in
the real world closely enough that the most important
processesare reasonablywell representedin such a model.
7%

Study

and Interpretation

of the Chemical

Characteristics

This is perhaps less likely for most other elements that


are subjectto redox reactionsowing either to unfavorable
kinetics or to lack of, or failure to include, some of the
fundamental thermodynamic data and reactions in the
model. Nevertheless,the pH-Eh diagram is a very useful
method of clarifying complicated chemical relationships
and illustrating boundary conditions and thermodynamically permissible reactions and products.
Two general types of information are furnished by
thesediagrams. If the total solute activities are specified,
the diagram can show pH-Eh domains in which various
solid specieswill be thermodynamically stable.Diagrams
stressingthis kind of information have sometimes been
called stability field diagrams.Conversely, if a a certain
assemblageof solids is specified, the diagram can be used
to show metal solubilities under a range of pH and Eh
conditions. Figures 14and 15illustrate thesetwo applications for the iron + water + oxygen + sulfur + carbon
dioxide system.
Becauseof the broad usefulnessof the diagrams, the
stepsin their preparation are given in some detail so that
the reader may apply the technique and understand its
implications.
For the calculations usedin constructing thesediagrams, standard conditions (2SC and 1 atmosphere)
were assumed,and the calculations are in terms of thermodynamic activities rather than concentrations.
The water-stability region obviously defines the
range of conditions to be expected at equilibrium in
aqueoussystems.Upper and lower limits for water stability in terms of Eh and pH are computed from the
relationship (all half-reactions are written asreductions):

2H+2e-=H:!(g).
Chemical thermodynamic data neededfor construction of figures 14and 15and referencesto sourcesof data
are given in tables 30 and 31 (appendix). From the
standard free energies, AGoa is computed for the first
equation (as stated in the discussion of electrochemical
equilibrium, the standard free energy of the electron
must be taken as zero):
2AG H~o(~)-AG~~(~)-~AG~~=AGR

-113.38-O-O=- 113.38kcal.
The kilocalorie units are converted to a standardpotential
by the relationship
E=--

-AGR

TZF
of Natural

Water

in which n is the number of electrons shown in the


half-reaction and F is the faraday constant, equal to
23.06 to give potentials in volts:
E=-

+113.38
=1.229 V.
4X23.06

-AGO,
2.303RT

=lw Keg.

The value of 2.303 R T under the specified conditions is


1.364. The value obtained for K eq is 10 17.Hence,

The Nernst equation (discussedin the section Electrochemical Equilibrium) applied to this reaction is
Eh=E+

2.303RT
nF

Again specifying that

log (Po,x[H+]).

[FeOH2]=[Fe3]
The value of 2.303RT/F for standard conditions is
0.0592. The maximum partial pressureof oxygen permissiblein the systemasdefined(1 atmospheretotal pressure)
is 1 atmosphere and n=4. The relationship thus reduces
to
Eh=1.229-0.0592

pH,

which defines the upper stability boundary for water.


The lower boundary similarly calculated is
Eh=O.OOO-0.0592
pH.
Boundaries between ferric and ferrous solute species
domains are fixed by calculations typified by that for
FeOH and Fe:
FeOH++H++e-=Fe2++Hz0.

The value of AGK is determined for this half-reaction as


indicated aboveand is usedto derive the standard potential E, which in this caseis equal to 0.898V. The Nernst
equation gives the value for Eh as follows:

Eh=O.898+0.0592log

leads to

which is the vertical boundary between the two domains.


These boundaries obviously terminate when they intersect. If the mathematics are done correctly, changesin
slope of the horizontal or sloping boundaries should
occur only where they intersect or are intersected by a
vertical boundary.
Lines designating stability domains of solids are
located by similar reasoning. For a boundary between
Fe(OH)s(c) and Fe2,

The standard potential for this half-reaction is 0.994 V


and
Eh=O.994+0.05921og

W+13

[Fe]

A value for [Fe] must be specified to fix this boundary.


For [Fe]= 10m6
I moles/L (56 pg/L),

[FeOH+] [H]
[Fe2]

Eh=1.349-0.178

At the boundary between the speciesdomains, [FeOH]


= [Fe]; hence the relationship becomes

pH.

Activities of solids and of HzO(l) are assignedvalues of


1.0throughout and do not appearin the Nernst equation.
The boundaries for some solids may involve activities of
dissolved sulfur species,for example,

Eh=O.898-0.0592pH.
Boundariesbetween speciesin which iron is at the same
oxidation state are fixed by mass-law equilibria. For
example,

or dissolved carbon dioxide species,


Fe2+HC03=FeC03(c)+Ht.

Fe3++HzO=FeOH2++H+.

A specified total concentration for sulfate was used in


drawing figure 14. The mass-law statement for siderite

The equilibrium constant is obtained from


Significance

of Properties

and Constituents

Reported

in Water

Analyses

79

precipitation must be supplemented with equilibrium


equations for carbon dioxide speciesto derive from the
total amount specified the fraction that is present as
HC03 at the pH of the boundary.
Thermodynamic data used in drawing figures 14
and 15 were mostly taken from Wagman and others
(1968) or from Robie and others (1978) and some differ
slightly from values used in the earlier edition of this
book. As noted earlier in this book, some writers have
preferred to usea notation pE (Sillen, 1967b, p. 52) or
per (Stumm and Morgan, 1981,p. 422-436) in place of
Eh. The rationale for this notation is primarily that it
simplifies calculations in which redox and other types of
chemical equilibria must be considered simultaneously.
In this writers opinion, the concept of pE asa measureof
molar activity of electrons tends to obscure the electrochemical factors involved in redox processesand the
greater mathematical simplicity is not very important in
an age of electronic computers. However, the value of
the diagrams for summarizing chemical relationships is
not affected by the choice of notation.
Figure 14 representsa system containing a constant
total amount of dissolved sulfur and carbon dioxide
speciesequivalent to 96 mg/L as SO:- and 61 mg/L as
HCOi. Boundariesare drawn for a dissolvediron activity
of 1o-6O0molar (56 pg/L). Solids indicated by the
shadedareaswould be thermodynamically stable in their
designateddomains. Boundaries between solute species
are not sensitive to specified dissolved iron activity, but
the domains of solid specieswill increasein area if more
dissolved iron is present.
The boundaries for sulfides and elemental iron
shown in figure 14 extend below the water stability
boundary and were extended to that area only to show
what conditions are required for thermodynamic stability
of elemental iron. These conditions include the absence
of liquid water.
Under the conditions specified for figure 14,siderite
(FeCOcl) saturation is not reached.Therefore, FeC03 is
not a stable phaseand it does not havea stability domain
in the diagram. The more stable ferric or mixed valence
oxyhydroxides (magnetite, goethite, hematite, etc.) were
not considered as likely equilibrium speciesin aqueous
systemsat low temperatures.
Figure 15 is an iron solubility diagram for which
most conditions are the same as those for figure 14. The
lower solubility of 10e7Omolar is equivalent to 5.6 pg/L
and the upper limit shown is lo- molar, equivalent to 56 mg/L. The contours are plotted asthe stability
boundaries for solids when the indicated activities of
dissolved iron are present.Calculations can be extended
to show higher and lower iron solubilities, but there is
little practical significance in doing so. As iron content
becomesgreater,the solubility of ferrousiron is controlled
by siderite precipitation rather than by FeO. The bound80

Study

and Interpretation

of the Chemical

Characteristics

ary between Fe0 and FeCOa in the presence of the


specified bicarbonate activity would be at pH 8.55. This
correspondsto an Fe2 activity of about 100 pg/L. The
solubility of iron in the siderite-controlled area is also a
function of the bicarbonate speciesactivity specified.
Figure 15 demonstratesthat there are two general
Eh-pH conditions under which iron solubility is very
low. One of theseis a condition of strong reduction in the
presenceof sulfur and coversa w,idepH range,within the
field of stability for pyrite. The second is a condition of
moderate oxidation above a pH of 5 and is in the ferric
hydroxide stability region. Between these regions, and
especially at low pH, iron is relatively soluble. The
high-solubility regions at high pH are outside the range
that is common in natural systems.
It is evidentfrom the solubiliry diagram that exposing
an equilibrated system to relatively small shifts in Eh or
pH can cause great changes in iron solubility. Thus,

0 60

.i

-0.60

Fe* +

--------I
t

FC?S

-060
t

Water

-1.00

re,j.cT

.l-

10

F;05Jy{

12

14

PPJ

Figure
Iron

14. Fields
as a function

pressure.
dioxide

Activity
species

/G/L.

of Natural

Water

of stability

for solid

of Eh and
of sulfur
61 mg/L

pH

and

a: 25C

species

96 mg/L

as HC03.,

and

dissolved
and

forms

of

1 atmosphere

as SOS-,
dissolved

carbon
iron

56

when pyrite is exposed to oxygenated water or ferric


hydroxide is in contact with reducing substances,iron
will tend to go into solution. It also is evident that if pH
and dissolved iron activity and the nature of the solid are
known, an equilibrium Eh of the system can be calculated, or at least a range of possible values can be
given.
Within the usual pH range of natural water (pH
5-9), the maintenance of an Eh below 0.20 and above
-0.10 V (or lower if sulfide is absent) can permit a
considerable ferrous iron concentration in equilibrium.
This goesa long way toward explaining the behavior of
iron in underground water, where a relatively low Eh
can be attained owing to chemical reactions that may
deplete dissolved oxygen. The reaction of oxygen with
pyrite is itself such a process,and the reaction products
SOi- and Fe2 may be transported stably in the moving
solution after the oxygen that might react with Fe to

-,ool
0

12

10

14

PH

Figure

1.5. Equilibrium

of Eh and
sulfur
61 mg/L

pH at 25C

species

96 mg/L

activity
and

of dissolved

1 atmosphere

as SO:-,

and

Iron

as a function

pressure.
carbon

dioxide

Actlvrty

of

species

as HCOs-.
Significance

convert it to Fe3 hasbeen usedup. It is notable also that


in a considerablepart of this rangeof Eh, the solubility of
iron may be controlled by precipitation of ferrous carbonate. In this part of the system, iron solubility is a function
of pH and dissolved bicarbonate speciesbut is independent of Eh. Thermodynamics of siderite control of ferrous
iron solubility were discussed by the writer in earlier
papers (Hem, 1960, 1965). If an iron-bearing ground
water of this type dissolvesoxygen from the air, however,
iron is oxidized to the ferric form, which may precipitate
as ferric hydroxide.
The positions of solubility contours like those in
figure 15 are strongly influenced by the thermodynamic
propertiesof the postulated solids. The free energy value
used for Fe(OH)a(c) in figure 15 is -166.0 kcal/mole,
which representsa poorly crystallized precipitate that
has undergonesome aging. According to Langmuir and
Whittemore (1971) natural fresh precipitates may have
JGo valuesashigh as-164.5 kcal/mole, but more stable
material is produced on aging. Attainment of significant
crystallinity may be very slow and, according to these
authors, could take thousandsof years if dissolved iron
activity is below 50 pg/L. More stable oxyhydroxides
such as hematite and goethite have very low solubilities,
but their roles in equilibrium control of iron solubility in
freshwater natural systemsprobably are seldom major.
Redox equilibrium calculations are useful in some
other aspectsof the aqueous chemistry of iron. Barnes
and Clarke (1969), for example applied such calculations
in studiesof the tendency of waters to corrode or encrust
steel well casingsand screens,and much of the impetus
for the development of pH-Eh diagrams by Pourbaix
was the need for a better theoretical understanding of
metal-corrosion processes.
Nevertheless,the field application of redox principles
and of pH-Eh diagramsis fraught with many uncertainties
and complexities, which can lead to erroneousinterpretations. Some errors stem from uncertainties about what is
producing the potential one might measurein a natural
water using a reference and inert metal electrode pair.
Such potentials will not be rigorously related to specific
dissolved metal ion activities if reversible equlibrium is
not attained. But nonequilibrium conditions are widely
prevalent in natural aqueoussystems.Some of the problems in interpreting such potentials were discussedby
Stumm and Morgan (1981, p. 490). Interpretations based
on measured potentials are likely to be very uncertain
unless the system can be well characterized by other
information.
Although the pH-potential diagram is an effective
meansof summarizingredox equilibria, the simplifications
that are required to prepare the diagram should be kept
in mind. Some of the limitations of the diagrams merit
attention.
Standard temperature and pressure (25C and 1
of Properties

and Constituents

Reported

in Water

Analyses

81

atmosphere) are generally specified. However, diagrams


could be prepared for other conditions if thermodynamic
data used are appropriate for those conditions. As a rule,
the departures from 2SC that are observed in natural
water systems, other than geothermal systems, do not
introduce enough deviation from standard conditions to
justify correcting for them. The effect of ionic strength
cannot be readily incorporated in the diagrams and
thermodynamic activities rather than concentrations must
be used.
A diagram such as figure 14 or 15 is an imperfect
representation of a real-world system in other important
ways. The prescribed constant total dissolved carbon
and sulfur species, for example, implies, incorrectly, that
one can evaluate the effect of changing some other
variable, such as pH, independently without concern
about its probable influence on total carbonate availability.
The mingling of open- and closed-system concepts that is
required for preparation of these diagrams limits their
applicability to natural conditions. The numerical value
of Eh, or pE, is an indicator of redox intensity, but it has
little value as an indicator of redox capacity. This constraint also applies in some degree to pH values.

Roles

of Bacteria

in Solution

and

Precipitation

of Iron

The requirement that rates of the reactions modeled


in pH-Eh diagrams be relatively fast has been noted.
Kinetics of relevant processes in the iron system have
been evaluated by many investigators.
Precipitation of ferric oxyhydroxide by oxygenation

The ways in which microbiota may influence the


occurrence of iron in water appear to be widely misunderstood. One result is a semi-folklore interpretation
that blames all high concentrations of iron in water on
activities of bacteria and leads to laborious and expensive
efforts to eliminate bacteria, and thus the troublesome
iron problem, by disinfecting wells, pipelines, and welldrilling equipment. These activities may have temporary
effects but do not eliminate all iron problems.
From a thermodynamic point of view, the ways in
which bacteria may be involved in the behavior of iron in
water include the following:
1. Processes in which bacteria exert a catalytic effect
to speed reactions that are thermodynamically
favorable but occur rather slowly in the absence
of bacteria. The biota may derive energy from
them but may not use the iron otherwise.
2. Processes that require a contribution of energy from
another source to alter the iron status and can be
promoted by bacteria that consume some other
substance as a source of energy.
Processes of the first type are involved in the oxidation of dissolved ferrous iron by Gallionella, Crenothrix,
and Leptothrix. These genera require oxygen; hence,
they live in environments in which ferrous iron is unstable.
The bacteria may become establis,hed in wells and remove
some iron by precipitation of ferric hydroxide before the
water gets to the surface of the ground. Sulfur-oxidizing
bacteria may exert an indirect effect on iron behavior by

of ferrous

catalyzing

Reaction

Rates

solutions

is rapid

at near-neutral

pH. Data

compiled by Sung and Morgan (1980) and determined


from their own investigations suggest that the process has
a half-time of about 18 minutes in aerated solutions with
ionic strengths of 0.02 or less at 25C and pH 6.84 in the
presence of about 550 mg/L HC03. The rate is very
sensitive to temperature and to OH- activity and above
about pH 5 (Stumm and Morgan, 1970, p. 538) is
second-order in [OK]. That is, the rate increases by a
factor of 100 for each unit increase in log [OH-]. Below
pH 3, the uncatalyzed rate of conversion of Fe to Fe3 is
very slow. The rate also is slowed by complexing of Fe,
especially by organic material (Theis and Singer, 1974).
Oxidation rates of pyrite and other sulfides are slow
in sterile systems but are greatly speeded by bacteria.
Biological mediation is generally necessary to reach equilibrium in sulfide systems. Measurements of rates and
evaluation of mechanisms, especially the oxidation of
pyrite by Fe3, were reported by Nordstrom, Jenne, and
Ball (1979).
Kinetic studies of ferric oxyhydroxide aging in relation to iron solubility were made by Langmuir and
Whittemore (1971).
a2

Study

and Interpretation

of the Chemical

Characteristics

reactions

that bring

iron

into

solution,

the

oxidation of pyrite, for example Bacterial catalysis may


increase the rate of conversion of Fe2 to Fe3 in acid
solutions in the presence of pyrite by as much as six
orders of magnitude and can be a major factor in generation of acid mine drainage (Nordstrom, Jenne, and Ball,
1979).
Processes of the second type typically involve ironand sulfur-reducing species which require an oxidizable
substance, usually organic, as a source of food and energy.
Reduction or oxidation of iron may occur incidentally,
and iron may play no essential role in the life processes of
the bacteria. Some species of bacteria that have been
shown to influence the precipitation of iron were discussed
by Clark and others (1967). There have been many
studies in microbiology concerning the roles of iron in
the lives of these microorganisms (Neilands, 1974).
The net effect of bacteria may be either to increase
or to decrease dissolved-iron concentrations in water
after it is intercepted by a well. B.acterial colonies in wells
and pipelines may be partly dislodged by moving water
from time to time and thus give rtse to occasional releases
of accumulated ferric hydroxide. The occurrence of iron
of Natural

Water

in ground water, however, is primarily a chemical phenomenon and cannot be ascribed solely to the bacteria.
The biota associated with iron solution and deposition
may aggravate some types of problems related to ironbearing water. These biota are abundant and will inevitably become established in favorable situations.
Different types of bacteria may live together in a
symbiotic relationship, with one species providing food
for another. Sulfur reducing and oxidizing species may
establish such relationships and greatly aggravate problems with well performance and corrosion of iron pipe
or well casing and other exposed metallic iron in watersupply systems.

Occurrence

of Iron

in Water

Water in a flowing surface stream that is fully


aerated should not contain more than a few micrograms
per liter of uncomplexed dissolved iron at equilibrium in
the pH range of about 6.5 to 8.5. The higher concentrations sometimes reported i-n such waters are generally
particulates (Kennedy and others, 1974) small enough to
pass through 0.45 pm porosity filter membranes. A
typical river-water concentration of IO pg/L is given in
analysis 8, table 12. The effect of organic complexing is
indicated by analysis 4, table 14, which is for a stream in
northeastern Minnesota. The reported color for this water
is 140 units, indicating a large amount of dissolved
organic material. The iron in solution, 1.4 mg/L, is
presumably complexed or stabilized by this material.
Another condition is shown in analysis 7, table 14,
which is for a Pennsylvania stream carrying drainage
from coal.mines. The water has a pH of 3.0 and 15 mg/L
of iron. Lower pH and higher iron concentration can
occur in coal-mine drainage water. The water represented
by analysis 6 in table 14, which is from a Michigan iron
mine, has a pH of 7.5 and an iron content of only 0.31
mg/L. Iron ores in the Lake Superior region are primarily
ferric oxides and silicates not readily soluble in water.
Analysis 3, table 18, represents water having a pH of 1.9
and 33 mg/L of iron.
In lakes and reservoirs in which a stratified condition
becomes established, water at and near the bottom may
become enriched in organic matter and depleted in oxygen
and may attain a low Eh. Ferrous iron can be retained in
solution in water of this type to the extent of many
milligrams per liter (Livingstone, 1963, p. G 11). The
iron content of lake water also can be influenced by
aquatic vegetation, both rooted and free-floating forms
(Oborn and Hem, 1962).
Occurrence of iron in ground waters of near-neutral
pH can generally be at least qualitatively explained by
chemical reactions that have been described and summarized in the pH-Eh diagrams. Deviations from these
Significance

simple models generally can be ascribed to water circulation and mixing mechanisms.
When it reaches the ground-water body, recharge
can be expected to contain oxygen acquired from exposure
to air. Dissolved-oxygen activity of IO- 5 moles/L represents approximate atmospheric saturation at ordinary
air temperatures. If this amount of oxygen is depleted by
reacting with pyrite, the final solution could contain
about 5 mg/L of ferrous iron and about 17 mg/L of
sulfate. A once-through continuous-flow system in which
oxygen is not replenished after this reaction has occurred
should have a final iron concentration in the water
approaching 5 mg/L. Higher iron concentrations in such
systems are common, however-perhaps
owing to faster
transport of oxygen than water in the unsaturated zone
or to occasional flow reversals accompanied by reactions
between ferric iron and pyrite.
Higher iron concentrations in simple hydraulic systems could result from interactions between oxidized
iron minerals and organic matter or by dissolution of
FeC03. Such waters should be relatively low in sulfate.
The latter effect may not be diagnostic, however, because
sulfate produced by pyrite oxidation may be reduced
again in some systems and lost from solution.
Analyses I, 2, and 3 in table 14 probably typify
iron-containing ground water where a pyrite-oxidation
mechanism is plausible. Regional development of a zone
of oxidation near the surface overlying a zone of reduction
at depth was described for interbedded sands of the
Mississippi Embayment in northeastern Texas by Broom
(1966). The ground water in this region is characterized
by low iron concentrations in the upper oxidized and
lower reduced strata, but iron concentrations are high
near the contact between zones at intermediate depth
where oxidation is currently active (Hem, 1965). Chemical zonations of iron in a lateral pattern in confined or
semiconfined aquifers in parts of New Jersey were described by Langmuir (1969a).
Ground water having a pH of between 6 and 8 can
be sufficiently reducing to retain as much as 50 mg/L of
ferrous iron at equilibrium, when bicarbonate activity
does not exceed 61 mg/L. In many areas, the occurrence
of 1.O- IO mg/L of iron in ground water is common. This
type of water is clear when first drawn from the well but
soon becomes cloudy and then brown from precipitating
ferric hydroxide. Wells that yield water of this type may
appear to be erratically distributed around an area, and
some exhibit changes in composition of their water with
time that are difficult for hydrologists to explain. For
example, in eastern Maryland and at other localities of
the Atlantic Coastal Plain, the permeability of some
coastal sedimentary beds is much greater than others,
and a well may encounter solutions with different oxidation-reduction potentials at different depths. Mingling of
these waters at contacts between strata and particularly
of Properties

and Constituents

Reported

in Water

Analyses

a3

sented by analyses 1-3, 8, and 9, table 14, had largely


been lost from solution before analysis. Concentrations
of iron reported are the totals originally present in a
filtered, acidified fraction of the sample. The oxidation
and precipitation reactions decreasedthe pH and altered
the alkalinity originally presentin the unacidified solutions
used for the rest of the analysis. Although it might be
possible to reconstruct these data to give the original
composition of the water at the time of sampling, it is not
common practice to do so. The effect of precipitation of
iron may be readily noticeable in dilute solutions such as
the one representedby analysis 9, table 14.

wherethe strataarelocally short-circuited,either naturally


or by a well that penetratesthem, can causedeposition or
solution of iron minerals. In many wells, corrosion of the
iron casing adds iron to the water pumped out, and iron
may be precipitated in various forms within the well
under some conditions.
Brines such as that representedby analysis 5, table
14, can retain iron that is stabilized by complex ion
formation.
Becauseof the unstable nature of ferrous iron in
samples of alkaline ground water, the iron orignially
present may have been oxidized and precipitated by the
time the analysis is made. To ensurethat concentrations
of iron reported in analyses represent the amounts in
solution at the time of sampling, waters suspected of
containing iron should be filtered at the time of sampling,
and an aliquot of the filtrate acidified to prevent precipitation. The iron in untreated portions of samples repre-

Manganese

Although manganeseis one of the more abundant


metallic elements, there is only about one-fiftieth as
Table 14. Analyses

[Analyses by U S Geological Survey. Date below sample number is date ofcollectton.


21), 8, Sampson (1929, p. 298)]

1
May 28, 1952

Constituent

mg/L

Mar.

meq/L

Silica (Sia)
............................
20
....................
Aluminum (Al) ..................................................................
Iron (Fe) ..................................
2.3
....................
Manganese (Mn) ......................
.oo
....................
Calcium (Ca) ............................
126
6.29
Magnestum (Mg) ......................
43
3.54
Sodium (Na) ............................
13
.56
Potassium (K) ___.______________._______2.1
.05
Bicarbonate (HCO:s) ................
440
7.21
Sulfate (S04) ............................
139
2.89
Chloride (Cl) ............................
8.0
23
Fluoride (F) ..............................
.7
.04
Nttrate (NO.,)
..........................
.2
.oo
Dissolved
Residue

............................
on evaporatton

.____

Hardness as CaCO,{ ..................


Noncarbonate

......................

conductance

(mtcromhos
pH

1.
2
3.
4.
5.

04

3
Feb. 27, 1952

meq/L

mg/L

12
....................
1.2
2.9
....................
...................................
2.7
,135
,164
2.0
35
1.522
1.7
,044
100
1.639
5.6
,117
2.0
,056
.I
,005
.6
,010

mg/L

meq/L

4
l-31,

mg/L

1962
meq/L

II

26
I.2
10
8.8
84
34
2.9
65
71
2.0
.3
.O

Oct.

1.4
.439
,691
1.478
,074
1.065
1.478
.056
,016
,000

I8
8.0

,898
,658

9.3

.40

69
29
64

1.131
,604
.181
_

2.9

,046

594
571
490

113

187
180
56
3
264

101

....................

15

I31

885

162

156
78
21
188

at 25C).

7.6
1

....................

..................................

__ ................

......................................

Color ........................................
Acidity

2
8, 1952

I. 5, and 6, U S. Geologtcal Survey, unpubhshed data. 2.3.

solids:

Calculated

Specific

Sources ofdata:

of

as HLSO,, (total)

7.4
23

....................

6.4
7

6.9
140

...................................

Well 3, Nelson Rd., Water Works, Columbus, Ohto. Depth, 117 ft, temperature, 13.3C. Water from glacral ontwash.
Well 798-50. pubhc supply, Memphrs, Term. Depth, 1,310 It, temperature, 22 2C. Water from sand of the Wdcox Formatron
Well 5.290-I. 6 mi southeast of Maryvdle, Blount County, Term. Depth, 66 fi; temperature, 14 4C Water from the Chattanooga Shale
Partrtdge Rover near Aurora. Mmn. Composrte sample. Mean drscharge, 30.8 cfs.
Brme produced wrth orl from well m NW l/4 see. 3, T. I I N , R 12 E , Okmulgee County. Okla Depth, 2,394 ft Water from the Grlcrease sand ofdrdlers. Atoka
Formation.

Study

and

Interpretation

of the

Chemical

Characteristics

of

Natural

Water

much manganesein the Earths crust as there is iron.


Manganeseis not an essential constituent of any of the
more common silicate rock minerals, but it can substitute
for iron, magnesium,or calcium in silicate structures.
The chemistry of manganeseis somewhat like that
of iron in that both metals participate in redox processes
in weathering environments. Manganese,however, has
three possiblevalencestatesin such environments rather
than two (2+, 3+, and 4+) and can form a wide variety of
mixed-valenceoxides. The 3+ speciesare unstablein that
they may disproportionate. That is, two Mn3 ions may
interact spontaneously to produce one Mr? and one
Mn4+ ion, and these products are more stable thermodynamically than the original Mn3 species.This ability
hasa number of interesting ramifications in the chemistry
of the element in natural water.
Manganese is an undesirable impurity in water
supplies, mainly owing to a tendency to deposit black
waters

containing

oxide stains.The recommendedupperlimit for manganese


in public water supplies in the United States is 0.05
mg/L (50 pg/L) (NAS-NAE, 1972). No mandatory
limit is specified for this element by the U.S. Environmental Protection Agency. It is an essential element for
both plant and animal life forms.
Sources

of Manganese

Many igneous and metamorphic minerals contain


divalent manganeseasa minor constituent. It is a significant constituent of basalt and many olivines and of
pyroxene and amphibole. Small amounts commonly are
present in dolomite and limestone, substituting for calcium. The silicate, rhodonite (MnSiOz), and the carbonate, rhodochrosite (MnCO:<), are pink to red minerals
sometimes used as gemstones.
When divalent manganeseis released to aqueous
solution during weathering, it is somewhat more stable

iron

and 9, Scott and Barker( 1962, p. 63, 101); 4, U.S. Geological Survey Water-Supply Paper 1948 (p 297); 7, U.S. GeologIcal Survey Water-Supply Paper 1022 (p

5
Mar. 11, 1952
wm

9.1

epm

. .._...............

6
Jan. 30, 1952
mg/L

8.1

32
7,470
1,330

43,800
129
76
47
83,800
0

meq/L

mg/L

21
____....__________._
29
15
10
13.17
119
1.40
68
2.26
17
.80 >
0
1.oo
817
15.76
22
68
.04
.I
.oo
.4

137,000
140,000

....................
....................

1,280
1,180

....................
.........................

1,260

24,200
24,100
146,000

....................
....................
....

730
679
1,460

....................

845
845
1,780

7.4
15

6.
7.
8
9.

7
Aug. 8, 1944

....................
....................

7.5
2

....................
....................

8
June 24, 1921

meq/L

mg/L

9
Oct. 26, 1954

meq/L

meq/L

_____.............._ 23
3.23
.81
.36
5.94
5.59
.I4

.oo
17.01
.62
.Ol
.Ol

7.9
_________________...
.6
.._...._.______........................
.._______________ 11
4.8
.32
.._......______._......................
136
6.19
8.4
35
2.88
1.5
1.5
960
41.74
3.6
249
4.08
30
1,260
26.23
5.9
734
20.70
1.8
..________._...................... .l
7.5
.12
.4

.................... 3,280

...............
.... 3,450
484
....................
....................
....................
280
....................
.. ........... ....... ................. ................

3.0
8
342

,419
,123
,065
,092
,492
,123
.05 1
.005
,006

47
44
27
2
68.8
6.3
3

Dramage at collar, drdl hole 89, 7th level, Mather A iron mine, Ishpemmg, Mlch Temperature. 15.1C.
Shamokin Creek at WeIghscale, Pa. Discharge, 64 2 cfs, affected by drainage from coal rnlws
Flowing well, Minneapolis, St. Paul, and Sault Ste. Marie R.R 1 Enderhn, Ramsom County, N. Dak Depth,
613ft
City well 4. Fulton, MISS.Depth, 210 ft; temperature, 17 2C Water from the Tuscaloosa Formation

Significance

mg/L

of Properties

and Constituents

waterfromtheDakota
sandstone,

Reported

in Water

Analyses

a5

toward oxidation than is ferrousiron. Generally, however,


if it is in contact with the atmosphere,it will be precipitated, at sites where pH is high enough, as a crust of
manganese4+ oxide. Theseencrustations generally contain a substantial quantity of coprecipitated iron and,
under some conditions, significant amounts of other
metal ions as well, especially cobalt, lead, zinc, copper,
nickel and barium. A particularly favorable substrate for
manganeseoxide precipitation is a previously existing
manganeseoxide surface.Thus the deposit may become
thicker with time and form nodules around some central
nucleus, as on the bottoms of certain lakes and at many
localities on the ocean bottom. Small, discrete particles
of oxide or coatings on other particles are widely distributed in stream sedimentsand soils. Coatings of manganeseoxides in streambedsoccur in many places.These
coatings generally do not become very thick, owing to
mechanical erosion and changesin water properties. In
someareasmanganeseoxide may accumulate with other
material in the form of a bog. The depositsin streamsand
bogs in New Brunswick, Canada, were described by
Brown (1964). Other descriptions of streambed deposits
concern streams in Sweden (Ljunggren, 1952, 1953,
1955),Colorado (Theobald and others, 1963) and Maine
(Canneyand Post, 1964;Nowlan, 1976).The mechanisms
that produce these deposits are different. The deposits
studied by Ljunggren seemedto be associated with an
aquatic moss.In the other areas,however, the deposition
seemedto occur without any organic intervention. Moran
and Wentz (1974) and Wentz (1974) studied Colorado
mountain streamsaffectedby drainagefrom mines.There,
as in the area described by Theobald and others, a
low-pH,

metal-enriched

water is neutralized

by mixing

Study

and

Interpretation

of the

Chemical

Characteristics

Form of Dissolved

Manganese

Under conditions to be expected in natural-water


systems, any dissolved manganese will be in the 2+
oxidation state. The ion Mr?+ will predominate in most
situations. The hydroxide complex MnOH becomesthe
principal form above pH 10.5,and anionic forms will be
significant at pH 12.0 or higher. The complex ion
MnHC03 can be important in solutions having bicarbonate concentrations near 1,000 mg/L as HCOa (Hem,
1963a,b). The MnS04(aq) ion pair could be important
in solutions in which sulfate activity is greater than a few
hundred milligrams per liter, judging from stability data
reported by Nair and Nancollas (1959). Organic complexes of manganese

may play a significant

role in its

transport in some situations. In general, however, the


Mn2+ion is considerably more stablethan Fe in aerated
water, and it can be transported at higher concentrations
without the protection of complexation.
Some Mn speciesmay be stable in strongly acid
solutions, and stable organic complexes of Mn3+ are
known to exist under some conditions. However, the
role of Mn in natural systemsprobably is related more
closely to the instability of this form of the element with
respectto disproportionation.
Manganeseat higher oxidation states than Mn4+is
not likely to occur to a significant extent in natural
aqueous systems. Very small amounts of such species
could theoretically occur at very high pH. Permanganate
(Mn04-) is used extensively in water treatment as an
oxidant for removal of iron, manganese,and organic
material. The Mn02 produced in such interactions is
removed by filtration.

with normal stream water, and much of the dissolvedmetal load is precipitated.
Manganese nodules in Oneida Lake, N.Y., were
describedby Dean and Greeson(1979). Literature on all
aspectsof marine manganeseoxide nodulesis voluminous
and growing.
In lakes and reservoirs where thermal stratification
develops,the bottom sedimentsmay become anoxic and
manganeseoxide previously deposited may be reduced
and dissolved. Water drawn from the deeper part of the
supply reservoir may at times contain significant concentrations of dissolvedMr?. Reductionof manganeseoxides
at depth in buried sediment may be a factor in accumulations of manganesenodules in many environments. Diffusing Mn can be redepositedas oxide when it reaches
the surfaceof the sediment layer and encountersoxygenated water (Callender and Bowser, 1980).
Manganeseis an essentialelement in plant metabolism, and it is to be expected that organic circulation of
manganesecan influence its occurrence in natural water.
Specific mention of manganese accumulation in tree
leavesappears in published literature. Some species of
86

treesare much more effective accumulators than others.


Bless and Steiner (1960) found considerable quantities
in the leavesof the chestnut oak.,and Ljunggren (1951)
reported similar findings for the needlesof sprucetreesin
southern Sweden. Aquatic plants were noted by Oborn
(1964) to be accumulators of manganese.Manganesein
plant parts that die back or are shed becomes available
for solution in runoff and soil moisture. The importance
of this source of manganesein river water is not completely known, although some preliminary studies by
Slack and Feltz (1968) of the effects of fallen leaveson
the water quality of a small stream in Virginia showed it
could be important at times.
Lateritic weathering processeshave produced manganeseoxide accumulations of economic importance in
many locations, and manganeseoxides are a constituent
of the dark stain (desert varnish) present on rocks in
arid regions (Hem, 1964, p. 10).

Quantification

of Manganese

Redox

Processes

Equilibria among dissovedand solid forms of manganesein the 2+, 3+, and 4+ oxidation states can be
of

Natural

Water

conveniently summarized in an pH-Eh diagram (fig. 16).


Use of the diagram to explain natural aqueoussystems,
however, is not as simple and straightforward as in the
caseof iron. The two pH-Eh diagrams for iron specify
oxyhydroxides that are not the most stablespeciesknown.
However, natural systems appear to be well enough
controlled by equilibria involving thesemetastablespecies
to justify their use in the model.
For manganese,the situation is less clear cut. The
common, naturally occurring oxides are generally forms
of MnOa, when the manganeseis at the Mn oxidation
state, or mixed-valence oxides in which the oxidation
state of manganeseis less than +4.0 but substantially
above +3.0. Specieswhere the oxidation state is +3.0 as
in MnOOH or below 3.0 as in Mns04 are readily producedwhen Mn2 solutesare oxidized by air in laboratory
experiments. Although theseoxides do occur naturally,
they are relatively uncommon.

Water

oxldlred

gives an equilibrium activity of dissolved Mn2 of less


than 1 nanogram/L (ng/L) at pH 7.00 in aerated water.
Apparently this simple equilibrium mechanism does not
control manganeseactivity in such systems.
A chemical mechanismfor manganeseoxidation in
aerated water that may help explain the generally observedbehavior of this element hasbeenproposed(Hem,
1978). In this model, precipitation of the oxide occurs in
an open system and manganesewill ultimately attain a
steady-stateconcentation that is a function of kinetics of
oxidation and of disproportionation of an Mn3+ intermediate. Essentially, this mechanismproposesa first step
in which an oxide containing Mn3 is produced and a
second step immediately following in which two of the
Mn3 ions still at the solid-solution interface disproportionate. The Mn4+ ion that is formed enters the oxide
crystalstructureand the Mn2 ion againbecomesavailable
for interaction with Oz(aq).
The coupled oxidation-disproportionation mechanism can be summarized graphically by plotting equilibrium lines for the reactions, for example,

Mn

0 60

manganese

-0 60

Water

Thermodynamic data used for preparing figure 16


are given in tables 30 and 32. If certain other forms of
oxyhydroxides or other published stability data are used,
the boundaries in the figure would be shifted somewhat,
but the general features would remain about the same.
The total activities of sulfur and carbon dioxide species
are 10m3
Omolar throughout, as in figure 14 and 15. The
manganoussulfide,alabandite,doesnot haveasimportant
an effectasthe iron sulfides,but the manganesecarbonate,
rhodochrosite, appearsto have more influence than the
correspondingferrouscarbonate,siderite.Figure 16shows
solid stability domains and equilibrium manganeseactivities from 1O-3ooto 1O-7oomoles/L (55 mg/L to 5.5
/a/L).
Natural oxygenatedwater commonly contains manganeseconcentrations near those predicted by figure 16
for equilibria involving yMnOOH but does not usually
havethe extremelysmall concentrationsthat are permitted
in much of the MnO2 stability field. For example, the
chemical reaction

reduced

1. 3Mn2+%02(aq)+3HzO=MnaO~(c)+6H

-0 60

-lOOt,,,,,,,,,,,,,l
4

and
10

2. Mn304(c)+4Ht=MnOz(c)+2HzO+2Mnzt

14

12

PH

Figure 16. Fields


librium
and
sulfur

dissolved
pH

at 25OC

species

61 mg/L

of stability
manganese
and

96 mg/Las

of manganese

solids

activity

as a function

and

equiof Eh

1 atmosphere
pressure.
Activity
of
and carbon
dioxide
species
SO,,

as HC03.

Significance

on a pH versuslog [Mn] grid (fig. 17).


In the region between lines 1 and 2 in figure 17,
both reactions are thermodynamically favored to go to
the right as written. If the process occurs in an open
of Properties

and Constituents

Reported

in Water

Analyses

87

system, that is, a system in which dissolved oxygen is


brought in by flowing water, and if there are continuing
suppliesof Mr? and continuing removal of surplus H, a
steady-statecontrol of Mt? concentration can be maintained. Figure 17showsthat for the solid speciesspecified,
activity of Mnzt could range from about 1Om6
to 10
molar at pH 7.5 (about 34 to about 0.3 /*g/L).
If a less stable Mn oxide is the initial form, the
positions of the lines would be shifted somewhat to the
right and solubilities would be greater, but the general
range of concentrations observed in oxygenated natural
waters are in reasonableaccord with values this precipitation mechanism forecasts.
Information available on the sources of samples
whose analyses are in tables 12-20 is not adequate to
evaluatethe applicability of this model closely. However,
7 of the 13 analyses that include specific manganese
concentrations and pH valuesare plotted in figure 17. Of
these, 6 plot between lines 1 and 2 and the other is just
outside line 2. Four of the remaining 6 analyses not
shown in figure 17 represent water having a pH of 4.0 or
less that clearly should not be expected to precipitate
manganeseoxide. The other 2 are substantially higher in
manganesethan the oxidation model would predict.

~too-J----l
70

60

50

40

so

SO

One of these(analysis 2, table 13) probably representsa


reduced system, and both it and the other (analysis 2,
table 18) have about the manganeseactivity one would
predict for equilibrium with rhodochrosite. Theseevaluations take into account effects of ionic strength and
complex-ion formation but are inexact becauseof uncertainties in the applicability of the laboratory pH determinations.
Researchby Morgan (1967), Wilson (1980) and
Sung and Morgan (1981) on kinetics of manganeseoxidation hasdemonstratedthe importance of solid surfaces
assitesfor oxidation and precipitation. Manganeseoxide
surfacesare especiallyfavorablefor catalysis of the oxidation process.Thesesurfaceshavea high density of negative
charge sites and are capable of adsorbing large numbers
of cations per unit area, when conditions favor this
process(Murray, 1974).
Whether the processesof adsorption or of coprecipitation arethe more important dependson the composition
of water in contact with the manganese oxides. The
oxides, in any event, are effective scavengers(Jenne,
1968,1975) and may be useful asindicators in geochemical prospecting for ore deposits (Chao and Anderson,
1974; Nowlan, 1976).
The rate of manganeseoxidation in aeratedsolutions
displays a second-orderdependenceon OH- activity and
is relatively slow in laboratory sy:stemsbelow pH 9.0. In
river waters in which favorable surfacesare plentiful, the
kinetic behavior may be pseudo-first-order; half-times of
the order of a few dayswere observedin the Susquehanna
River in Pennsylvania(Lewis, 1976) at near-neutral pH
and temperaturesof 5 to 20C.
The rate of oxidation also is strongly temperature
dependent, and lower temperatures (near OC) seem to
favor the formation of more fully oxidized species(Hem,
1981; Hem and others, 1982) as initial precipitates
(PMnOOH rather than Mn30d).
Bacteria may influence the rates of manganeseoxidation, probably in a manner similar to the effects of
bacteria in oxidation of ferrous iron. Tyler and Marshall
(1967) found Hyphomicrobium in depositsof manganese
oxide in pipelines. Schweisfurth hasconducted extensive
researchon manganeseoxidizing bacteria and has published many papers in Germany (for example, Schweisfurth, 1973; Jung and Schweisfurth, 1976).

100

Occurence
Figure

17. Activity

at 2SC

and

aerated

water.

Line

Line

Plotted

13-16

and

points

20 (e.g.,
Study

manganese

pressure

1, oxidation

2, disproportionation

Mr?.

00

of dissolved

1 atmosphere

of Mn

of MnaOl
represent

14-9=table

and Interpretation

at equilibrium

as a function
to form

to form

analytical

of pH

data

14, analysis
of the Chemical

MnaOd.

yh;lnOz
from

in

and
tables

9).
Characteristics

of Manganese

in Water

Manganeseis often present to the extent of more


than 1 mg/L in streamsthat have received acid drainage
from coal mines. Manganeseusually persistsin the water
for greater distances downstream from the pollution
source than the iron contained in the drainage inflows.
of Natural

Water

As the acidity is gradually neutralized, ferric hydroxide


precipitates first. Manganese,however, also disappears
from solution after a longer time. These effects can be
notedfor streamsin the Ohio River basin(U.S. Geological
Survey Water-Supply Paper 1948). Analysis 4, table 13,
presents water from a river influenced by acid mine
drainage from which most of the iron has been precipitated. Analysis 7, table 14, is for water affected by mine
drainagein which much iron was still present.Manganese
concentrations exceeding 1 mg/L have been measured
in water from small streams in the Northern United
Statesduring winter low-flow conditions (U.S. Geological
Survey, 1969, p. 240, for example). During long periods
of ice cover the dissolved-oxygen concentration in the
water is depleted and manganesemay be extracted from
organic-rich sediment. Recordsof composition of many
U.S. streamscollected since about 1970 include manganeseconcentrations. Thesedata indicate that concentrations up to a few hundred micrograms per liter occur in
many streamsat times. Some of the manganesereported
in these analyses may be particulate material that can
pass through 0.45 pm filter pores. Data obtained by
Kennedy and others (1974), however, suggestthat this is
a much lessimportant effect for manganesethan it is for
iron or aluminum.
Ground waters may contain more than 1.Omg/L of
manganeseunder somecircumstances.High iron concentrations may accompany the manganese,but this is not
invariably true. Analysis 1, table 13, is for a low-pH
water that is very high in manganese,but some of the
iron that might originally have been in solution in the
water could have been lost by precipitation of ferric
hydroxide in the sample after collection.
Analysis 2, table 13,representswater pumped from
two radial collector wells located along the Ohio River
at Parkersburg,W. Va. The manganeseconcentration of
this water is 1.3 mg/L, compared with only 40 r.lg/L of
iron. Excessivemanganeseconcentrationsalso havebeen
reported in other localities where wells were located
with the aim of withdrawing water from gravel deposits
adjacentto or within streamchannels.Streambedmaterials may include sandgrainsand pebbleshaving manganese
oxide coatings. The coatings may be reduced and dissolved by water of different composition that reaches
them. Dissolved organic solutes,for example, may reduce
the oxide and liberate Mn2. Anaysis 5 in table 13 and
analysis 9 in table 14 represent ground waters high in
iron that also contain several hundred micrograms per
liter of manganese.The manganesemay well have come
from the same initial source as the iron.
Many of the ground waters reported to carry large
manganeseconcentrations are from thermal springs.
White and others(1963, p. F50) reported someexamples
of these.In many places thesesprings seemto be closely
associated with manganeseoxide deposits. Analysis 3,
Significance

table 13, represents water from a thermal spring that


contains 0.31 mg/L of manganese.

Calcium

Calcium is the most abundant of the alkaline-earth


metals and is a major constituent of many common rock
minerals. It is an essentialelement for plant and animal
life forms and is a major component of the solutes in
most natural water. Calcium has only one oxidation
state, Ca. Its behavior in natural aqueous systems is
generallygovernedby the availability of the more soluble
calcium-containing solids and by solution- and gas-phase
equilibria that involve carbon dioxide species,or by the
availability of sulfur in the form of sulfate. Calcium also
participates in cation-exchangeequilibria at aluminosilicate and other mineral surfaces.Solubility equilibrium
models have been used widely in studying the chemical
behavior of calcium.

Sources

of Calcium

Calcium is an essentialconstituent of many igneousrock minerals, especially of the chain silicates pyroxene
and amphibole,and the feldspars.The plagioclasefeldspar
group of minerals representsmixtures in various proportions of the end membersalbite, NaAlSi308, and anorthite,
CaAhSizOs. Calcium alsooccursin other silicate minerals
that are produced in metamorphism. Some calcium is,
therefore, to be expected in water that has been in
contact with igneous and metamorphic rock. The concentration generally is low, however, mainly becausethe
rate of decomposition of most igneous-rock minerals is
slow. The decomposition of anorthite can be represented
as

The normal composition of plagioclase feldspar lies between the pure sodium and pure calcium forms, and
decomposition will, therefore, generally yield both calcium and sodium and some soluble silica. Under some
conditions the solution may attain saturation with respect
to calcium carbonate.
The most common forms of calcium in sedimentary
rock are carbonates. The two crystalline forms, calcite
and aragonite, both have the formula CaC03, and the
mineral dolomite can be represented as CaMg(COs)z.
Limestone consists mostly of calcite with admixtures of
magnesiumcarbonateand other impurities. A carbonate
rock is commonly termed dolomite if the magnesium
of Properties

and Constituents

Reported

in Water

Analyses

89

is present in amounts approaching the theoretical 1 : 1


mole ratio with calcium. Other calcium mineralscommon
in sedimentsinclude the sulfatesgypsum (CaS04. 2H20)
and anhydrite (CaSO& and, more rarely, the fluoride,
fluorite (CaFa). Calcium is also a component of some
types of zeolites and montmorillonite.
In sandstoneand other detrital rocks, calcium carbonatecommonly is presentasa cement betweenparticles
or a partial filling of interstices.Calcium also is presentin
the form of adsorbedions on negatively charged mineral
surfacesin soils and rocks. Becausedivalent ions are held
more strongly than monovalent ions at surface charge
sites, and because calcium is generally the dominant
divalent ion in solution, most such sites are occupied by
calcium ions in the usual river or ground-water system.
Solute

Species

Calcium ions are rather large, having an ionic radius


near 1 angstrom. The charged field around the ion is
therefore not as intense as the fields of smaller divalent
ions. Calcium ions have a less strongly retained shell of
oriented water molecules surrounding them in solution.
The usual dissolved form can be simply represented as
the ion Cazt. Calcium does form complexes with some
organic anions, but the influence of such specieson the
concentrations of calcium in natural water is probably
not important. In some solutions, complexes such as
CaHCOi can exist. Data published by Greenwald (1941)
show that about 10 percent of the calcium might be in
this form if the bicarbonate concentration were near
1,000 mg/L. The ion pair CaSOd(aq)is more important.
In solutions in which sulfate concentrations exceed 1,000
mg/L, more than half the calcium could be presentin the
form of the CaS04 ion pair. Both of thesegeneralizations
assumethat the calcium concentration is small compared
with the bicarbonate or sulfate concentrations.
Garrels and Christ (1964, p. 96) also mentioned
hydroxide and carbonate ion pairs with calcium. These
speciescould be presentin appreciableconcentrations in
strongly alkaline solutions. Other calcium ion pairs, such
as those with phosphate,are known but are not likely to
influence the behavior of calcium in natural water to a
significant extent. However, small amounts of phosphate
are sometimesaddedin water treatment to inhibit CaCOs
precipitation and phosphate is a constituent of many
types of wastewater.
Chemical

Controls

of Calcium

Concentration

Equilibria involving carbonatesare the major factor


in limiting the solubility of calcium in most natural
water. The calcite dissolution-precipitation and dissolved
carbon dioxide speciesequilibria were introduced earlier
to demonstrate chemical equilibrium calculations, and
90

Study

and Interpretation

of the Chemical

Characteristics

the discussionof pH. Calculations required to determine


whether a particular solution ma.ybe in thermodynamic
equilibrium with calcite require values for activities of
Ca, HCOs- and H and solution temperature.Analytical
concentrationsmust be correctedfor ionic strengtheffects
and complexing, and equilibrium constantsusedmust be
appropriate for the temperature of the system.
Computer programs such as WATEQ (Truesdell
and Jones, 1974)can perform all the necessarycomputations. For many interim or preliminary studies of water
chemistry data, however,it may be convenient to evaluate
the calcite equilibrium by a graphical procedure. Two
forms of the calcite-solubility rel,ationshipare given here,
as plate 2a and 2b and in figure 18. Both entail some
simplifying assumptionsthat limit their applicability range
to some extent.
Plate 2 is for a system lacking a gas phase and is
based on the calcite-solubility equilibrium equation already extensively discussed:
[Ca] [HCO;]
[H+l -=c
The symbol K, representsthe solubdity equilibrium constant for calcite
in the reactlon CaCO,s+H*=HC03-+Ca

The illustration consists of two parts, a log-log grid of


calcium concentration versusbicarbonate concentration
(both in mg/L) and a pH-grid overlay. When in use, the
overlay is alined on the log-log plot so that the x-axes
match.
Equilibrium constantsat varioustemperatures,based
on data in published literature, are given in table 33. The
equilibrium solubility of calcite hasa strong temperature
dependencein this range. The solubility at 0 is more
than five times asgreatasat 50. To determineequilibrium
conditions for a single temperature in this range when
the effect of ion activity corrections can be ignored (ionic
strength -O.O), the overlay may be moved right or left
until the extension of the pH 8.0 line intersects the
appropriate point on the temperature scaleof the log-log
basesheet.A saturation index for a water sample having
this temperature can be obtained by plotting the point of
intersectionofthe calcium and bicarbonateconcentrations
given in the analysis and reading the equilibrium pH
from the pH grid. This pH is subtractedfrom the measured
pH given in the analysis to give the S.I. value. A positive
value for S.I. indicates supersaturation.
To permit the use of the graph for ionic strengths
>O.O, conversion of stoichiometric concentrations to
thermodynamic activities was incorporated in plate 2 by
including activity coefficients in the equilibrium constant.
The mass-lawexpressionin terms of concentrations, C,,,
may be written
of Natural

Water

CCa

%I~+

x CHC03-

cCaz+ x CHC03-

WI

-iHC03-

=K,

KS

= YC: YHC&

The activity coefficients (gamma terms) were evaluated


by meansof the Davies equation for values oflfrom 0.0
to 0.5. This relationship is
dI
-log yi = 0.5092,z 1iJI

- 0.21,

where 2; is the ionic charge and Z is ionic strength


(Butler, 1964, p. 473).
The ionic-strength scale on the pH-grid overlay
representsthe positions to which the pH 8.0 line would
be shifted by increased ionic strength. Simultaneous
calibration of the graphfor temperatureand ionic-strength
effects is accomplished by positioning the pH grid at the
point where the temperature of the solution matches its
ionic strength, using the scaleson the match line in the
lower parts of the two sheets.
Plate2 providesa simple meansof checkingchemical
analysesfor possibleconformance to calcite equilibrium.
The only computation required is for ionic strength, and
this can be obtained by using plate 1; as noted earlier, it
also may be possible to estimate a usable number from
the specific conductance. The graph does have some
practical and theoretical limitations, however, and it is
not applicable to all natural waters. Chemical analyses
report calcium asa total Ca*+,which includesany calcium
complexesor ion pairs that are present.Any equilibrium
calculation, however, must be basedon the actual species
of the solutes involved. Calcium forms complexes with
bicarbonate and sulfate. The calcium bicarbonate complexesdo not affect the usefulnessof plate 2 becausethe
basic data of Jacobson and Langmuir (1974) that were
used compensate for these complexes. However, the
effect of the calcium sulfate ion pair can be substantial,
and waters that contain more than a few hundred milligrams of sulfate per liter require a more comprehensive
mathematical treatment. Determinations of HC03 by
titration aresubjectto uncertaintiesnoted in the discussion
of alkalinity in a later sectionof this book. The bicarbonate
concentration required when using plate 2 is that fraction
of the alkalinity present as HC03 (not total alkalinity as
HC03).
The Daviesequation doesnot include a temperature
effect on the ion activity. Such a effect is indicated in the
extended Debye-Htickel expression used elsewhere in
Significance

this book. The designof plate 2 of necessityrequires that


temperature doesnot affect the activity coefficient calculation. This simplification probably does not introduce
important errors compared with those resulting from
other factors that are inherent in applying simple equilibrium models, however.
The systempostulated doesnot contain a gasphase,
and an equilibrium among dissolved CO2 species is
assumed.Measurementsof pH and bicarbonate concentration must be as closely representative of the actual
systemas possible.This generally requires measurement
of both properties in the field at the sampling site, at the
temperature of the water source (Barnes, 1964). The
calcium concentration in a sample can be stabilized by
acidifying the sample at the time of collection so that this
property can be measuredlater in the laboratory. Enough
additional determinations must be madeto permit calculation of ionic strength.
A lessreadily controllable requirement for applying
calcite equilibrium calculations is that the solid involved
is pure calcite and the reaction is directly reversible. The
calcium carbonate in most limestones has significant
amounts of other divalent cations substitued for Ca* in
the lattice. A congruent reversible equilibrium may not
be attainable in such a system,especially if the attacking
solution has a Ca:Mg ratio substantially different from
the solution in which the initial limestone was produced.
However, in many if not most practical systems,a relatively pure secondarycalcite precipitate can exist, and its
behavior may indeed approximate reversibility well
enoughto make the model useful. In some environments
films of organic matter may form on calcium carbonate
surfacesand inhibit precipitation or dissolution.
Plummer and Busenberg(1982) have provided additional solubility data for calcium carbonates that can
be usedover the temperaturerangeO-90C. A saturation
index (S.I.) for a water can be derived using plate 2 as
noted above. Most investigators seem to have thought
that experimental and other uncertainties in S.I. values
are unlikely to be lessthan kO.10 (Langmuir, 1971) and
any systemin which the graphically estimated S.I. value
is in the range -to.3 is probably close to equilibrium.
The determination of S.I. is useful asan indicator of
chemical stability and of probable behavior of ground
water encounteredin wells, where there may be concern
over possible deposition of carbonate precipitates on
well screens,gravel packs,and water-yielding rock faces.
Some aspects of this were discussed by Barnes and
Clarke (1969). Such precipitates may be mixtures of
various solids and can be expected if waters of differing
chemical composition enter the well during operation of
the pump.
The occurrence of siderite (FeCOa) in carbonate
precipitates that form in pipelines where some iron is
available either from the water itself or the distribution
of Properties

and Constituents

Reported

in Water

Analyses

91

The activity of dissolved calcium in such a system is


therefore specified if a value for PCO,is specified. The
equations required for computing the activity of Ca2
include five chemical equilibria and a cation-anion balanceequation, and enough information to permit calculating activity coefficients. Temperature effects can be
included by usingappropriateequilibrium constantsfrom
table 33. The final product could be a graph of calcium
concentration versusPCO,for specified temperature and
ionic strength. Calculations of this type were described
by Garrels and Christ (1964, p. 81-83).
Natural systems contain other solutes and solids,
and this kind of calculation has a somewhat limited
practical applicability. Figure 18 is a graph of calcite
solubility in terms of calcium concentration and CO2
partial pressurebasedon laboratory experimental data at
25C that were published by Frear and Johnston (1929).
The CO2 content of normal air is 0.03 percent (by
volume), or 0.0003 atmosphere.At this Pco2,the solubility indicated by the Frear and Johnston data in figure 18
for calcium in water in contact with air is about 20
mg/L. Garrels and Christ (1964, p. 83), using other

system was noted by Sontheimer and others (1981).


These authors suggestedthat siderite formation was an
essentialstep in the development of a protective calcium
carbonate coating, which prevents corrosion of waterdistribution pipes.
In the presence of a gas phase containing COz,
computations of calcite solubility can be made in terms
of the partial pressure of CO, and pH. This requires
consideration of carbon dioxide solubility, from Henrys
Law
P32CO31 =K
h
pC@2

and the first dissociation of carbonic acid

WCW Wl

=K,.

l32CO31

The activity of HzC03 is assumedto include any uncombined dissolvedCOz as well, in accord with conventional
practice. Combining these equations with the one for
calcite solubility gives, when all three are at equilibrium:
[Ca] K&pcq

50 -

=I.&.

Wl-

lo-

Valuesof the equilibrium constants from 0 to 50C are


given in table 33.
This expression gives an indication of the COz
content of the gas phase that provided a water with its
capacity

for dissolving

calcite. It might be interpreted

5-

for

ground water asindicative of the partial pressureof CO2


in the unsaturated zone through which recharge passed
asit moved toward the water table. Subsequentreaction
between CO2 speciesand calcite would presumably have
occurred in the absenceof a gasphase,and no replenishment of CO2 could occur. Depending on the nature of
water circulation and water table fluctuations, this may
or may not be a valid assumption. Some replenishment
of the dissolved CO2 in the zone of saturation might
occur in some systems,owing to SO,- reduction.
Observationsand calculations were madeby Deines
and others (1974), on the basis of 13Ccontents and the
foregoing equilibria, on ground-water systems in limestone in the Nittany Valley of Pennsylvania. The Pco,
valuesfor thesewaters rangedfrom 10-l 3to 10-l atmospheres.It appearedthat after equilibrating with the gas
phase the water generally reacted with calcite under
closed-systemconditions like those postulated for plate

:5

lo-

I
%

05-

fi

Ol-

8h

005-

g
2E

OOl-

0 005 -

a
F
:

OOOl-

0 0005 -

//
0.0001

0 00001

2.

L
10

SOLUBILITY

In the discussionsof pH and of the phaserule it was


shown that a closed system containing water, carbon
dioxide gas,and calcite would have 1 degreeof freedom.
92

0.00005 -

Study

and Interpretation

of the Chemical

Characteristics

Figure
at 25C
of Natural

18. Solubility
as a function
Water

4111,111

I11111

100
EXPRESSED

AS CALCIUM.

of calcium
of partial

carbonate
pressure

1000
IN MILLIGRAMS

(calcite)
of CO2.

PER LITER

in water

published thermodynamic data, calculated a solubility


of 16 mg/L under these conditions.
If the input of H is solely from the water itself and
carbon dioxide is absent, the solubility of calcite is 5.4
mg/L of calcium (Askew, 1923). It is of interest to note
that this solution would havea pH between 9.0 and 10.0.
Garrels and Christ (1964, p. 81) observed this experimentally. They further calculated (p. 88) that rainwater
that had dissolvedall the carbon dioxide possiblethrough
equilibrium with air and then was equilibrated with
calcite in the absenceof a gasphasecould dissolve hardly
any more calcium than pure water and would also reach
a high pH. Such a solution is not a very effective solvent
for calcite in a stoichiometric sense. However, other
solutes that may be present in rainwater (acid rain)
may substantially increaseits solvent power.
From figure 18 it is evident that the concentrations
of calcium that are common in waters in regions where
carbonate rocks occur generally are well above the level
specified for the partial pressureof carbon dioxide in the
atmosphere. Water percolating through the soil and the
unsaturated zone above the water table, however, is
exposedto air in pore spacesin the soil, which is greatly
enriched in carbon dioxide. Partial pressuresof carbon
dioxide in soil air are commonly lo-100 times the levels
reachedin the atmosphere(Bolt and Bruggenwert, 1978,
p. 11).The carbon dioxide content of soil air for the most
part results from plant respiration and from decay of
deadplant material and can be expectedto be greatestin
environments that support densestands of vegetation.
Observations and calculations of PC,, for groundwater systemscommonly give valuesbetween lo-. and
1O-25for PCO, in the gas phase. Values reported by
Deines and others (1974) already cited are typical of
systemsin temperate regions where vegetation is abundant. Other saucesof CO2 such as sulfate reduction and
oxidation of lignite fragments may be significant in some
systems.
For dolomitic terranes, a reversible equilibrium is
more difficult to attain, but a model postulating equilibrium has been shown by some workers to predict composition of ground water reasonably well (Barnes and
Back, 1964a;Langmuir, 1971). This equilibrium will be
considered more fully in discussing magnesium in the
next section.
In environmentsin which hydrogenions are supplied
for rock weathering by processesother than dissociation
of dissolved carbon dioxide species, for example, by
oxidation of sulfur or sulfides, calcium may be brought
into solution in amounts greater than the stoichiometric
equivalent of bicarbonate. In such a system, or where
water is in contact with solid gypsum or anhydrite, the
maximum calcium concentration that could be reached
would generally be determined by equilibria in which
gypsum is a stable solid. Systems in which calcium
Significance

solubility might be controlled by fluoride, phosphate,or


other anions likely to occur only at low concentrations
can be postulated but are likely to be rare in the real
world.
The effect of cation-exchangeprocesseson calcium
concentrationsin water differs from mineral dissolutions
or precipitations in that only cations are directly involved.
The requirement of anion-cation balance in the reacting
solution thus preventssimple cation exchangeor desorption from increasing or decreasingthe total ionic load.
Cation-exchange processescan bring about changes in
the ratio of calcium to other cations in solution. Also, this
effect in systems in which each liter of the water is in
contact with an extensive area of solid surface tends to
maintain a relatively constant Ca:Na ratio.
Reaction rates for calcium carbonate precipitation
have been studied rather extensively. The presenceof a
favorable solid surface is a substantial aid in starting
precipitation from a solution that is supersaturated.The
rate of attainment of calcite saturation in ground water
moving through a limestone aquifer depends on the
nature of the flow system and the intimacy of contact of
the water with calcite surfaces as well as on chemical
factors. Studies in Pennsylvania showed that water of
conduit type springswhere residencetime of the water
was short (days or weeks) commonly is unsaturated.
Water from diffuse type springs where residencetime
is measuredin months is generallynearsaturation (Deines
and others, 1974).
Studies by Plummer and Back (1980) showed that
a continuing irreversible dissolution of dolomite and
gypsumwith someprecipitation of calcite occurred along
flow paths in the Floridan aquifer in Florida, and in the
Madison limestone in South Dakota.
Occurrence

of Calcium

in Water

Applicability of the chemical models for calcium to


natural waters can be illustrated with some of the tabulated analysesin this paper.
Generally calcium is the predominant cation in
river water. Analysis 7 in table 15 is for the Cumberland
River at Smithland, Ky., near the point where it joins the
Ohio River. The pH of river water is influenced by many
factors,especiallyby photosynthesizingorganisms.These
organismscommonly causediurnal and seasonalfluctuations, with higher pH values on summer days and lower
values at night and during other periods of low photosynthetic activity. The measured pH in river water is
generally not well correlated with calcium and bicarbonate activities. However, the calcium concentration of the
Cumberland River water, 25 mg/L, is not greatly above
that which could occur in water in equilibrium with
calcite in contact with air (20 mg/L, asshown in fig. 18).
It is interesting to note that the averageconcentration of
of Properties

and Constituents

Reported

in Water

Analyses

93

on solid surfacesaround the edgesof the water body. A


conspicuous white deposit developed around the edges
of Lake Mead on the Colorado River between Arizona
and Nevada as a result of this process. In part, the
deposition in that reservoir also was related to inflows
that are high in calcium and to the solution of gypsum
bedsexposedin the reservoirin the first few decadesafter
the beginning of the storageof water (Howard, 1960, p.
119-124). A quantitative study of the changes that occurred in stored water composition, owing to gypsum
solution and calcium carbonate deposition, was made by
Bolke (1979) on the Flaming Gorge reservoir on the
Green River in Wyoming and IJtah. During the period
1963-75, there was a net gain of 1,947,OOOmetric tons
in dissolved-solidsload of the river attributable to effects
of this kind within the reservoir. The freshwater deposit
called marl, which is formed in many lakes, is made up
partly of calcium carbonate.

calcium in river water, shown as 13.4to 15 mg/L in table


3, is also not far from an equilibrium value for a system
containing calcite in contact with air. At equilibrium,
however, a pH much greater than is usually observed in
river water would be predicted.
Rivers in more arid regions, especially where some
of the more soluble rock types are exposed,tend to have
much higher dissolved-calcium concentrations. Analysis
8, table 15, is a discharge-weighted averageof the composition observedfor a year in the PecosRiver at Artesia
in southeasternNew Mexico. It shows clearly the influenceof gypsum, which is abundant in the Pecosdrainage
basin.
When river water is impounded in a storage reservoir, changesmay occur in calcium content as a result of
calcium carbonate precipitation. The increased pH near
the water surface, caused by algae and plankton, may
bring about supersaturation, and precipitation can occur

Table 15. Analyses

[Analyses by U.S Geological Survey. Date below sample number IS date of collection.

Mar.

Constituent

meq/L

mg/L

Nov.

meq/L

mg/L

which

Sources of data: 1, 4, and 9, Scott and Barker (1962, p. 19, 47, 59); 2,

2
May 20, 1950

1
28, 1952

of watersin

mg/L

3
25, 1949
meq/L

Mar.
mg/L

4
26, 1952
meq/L

130
....................
8.6
....................
22
....................
29
....................
41
....................
.05 ..........................................................................................................
Manganese (Mn) ...............................................................................................................................................................................................
93,500
4,665.65
636
31.74
2.395
144
7.19
Calcium (Ca) ............................
48
12,100
995.35
43
3.54
,296
55
4.52
Magnesium (Mg) ......................
3.6
28,100
1.222.35
Sodium (Na) ............................
17
.74
.091
29
1.24
2.1
299.17 1
Potassium (K) ..........................
1 11,700
)
0
.oo
143
2.34
2.491
622
Bicarbonate (HCOa) ................
152
10.19
17
.35
,067
1,570
32.69
3.2
60
1.25
Sulfate (Sod) ............................
255,000
7,193.55
24
.68
,226
53
1.49
Chloride (Cl) ............................
8.0
...................................................................................
.02
.4
.O
....................
Fluoride (F) ..............................
.29
..........................................
18
.3
.oo
Nitrate (N03) ......................................................................
Dissolved solids:
.................... 408,000
.................... 2,410
....................
670
Calculated ............................
148
....................

Silica (SiOa)

............................

Iron(Fe) ..................................

Residue

on evaporation

........................................................................................................................................

415,000

....................

.................... 1,760
...............................................................
586
Hardness as CaC03 ..................
135
....................
.................... 1,650
Noncarbonate ......................
....................
10
76
.................... 2,510
(3)
Specific conductance
269
.... 1,120
....................
....................
(micromhos at 25C).
5.29 ....................
7.5
..........................................................................................................
pH ............................................
Color .................................................................................................................................................................................................................
Includes strontium (Sr) 3,480 mg/L, 79.43 meq/L; bromide (Br) 3,720 mg/L,
Includes 0.70 mg/L lithium (Li) and 0.03 mg/L zinc (Zn).
aDensIty at 46% 1.275 g/mL.

46.56 meq/L;

iodide (I) 48 mg/L, 0.38 meq/L.

1. Big Spring, Huntsville, Ala Temperature, 16.1C. Water-bearing formation, Tuscumbla Limestone.
2 Spring on Havasu Creek near Grand Canyon, Ariz. Flow, 100 gpm; temperature, 19.4% Water-bearing formation, hmestone in the Supai Formation. Water
deposits travertine.
3. Jumping Springs, SE1/4 sec. 17, T. 26 S., R. 26 E., Eddy County, N. Mex. Flow, 5 gpm. Water-bearing formation, gypsum in the Castile FormatIon.
4. Brine well 3 Monroe, SE1/4 sec. 27, T. I4 N., R. 2 E , Midland, Mich. Depth, 5,150 ft; temperature, 46.1C. Water-bearing formation, Sylvama Sandstone.
94

Study

and

Interpretation

of the

Chemical

Characteristics

of

Natural

Water

In closed basins,water can escapeonly by evaporation, and the residual water can be expected to changein
composition after a fairly well defined evolutionary pattern. The lesssoluble substances,including calcium carbonate,arelost first, followed by calcium sulfate(Swenson
and Colby, 1955, p. 24-27). The Salton Seain Imperial
Valley, Calif., is fed by irrigation drainage, and since its
origin, during flooding by the Colorado River in 1907, it
has become rather highly mineralized. Hely and others
(1966) concluded that the water of the Salton Sea in
1965 was at saturation with respect to gypsum because
the water seemedincapable of dissolving powdered gypsum in a laboratory experiment.
Equilibrium solubilities can be expected to control
calcium concentrations in many ground-water systems.
Unfortunately, many published chemical analyses of
such waters are unsuitable for rigorous testing of calcitesolubility models owing to lack of accurate onsite measurementsfor pH and alkalinity. Most of the analysesin
calcium

is a major

6, and 10, US

mg/L

constituent

Geologtcal Survey, unpubhshed data; 3 and 5, Hendrickson

5
Jan. 26, 1948

6
June 23, 1949

meq/L

the tables in this book also have this shortcoming. However, if the acceptable uncertainty in saturation index
calculations is broadened,the application of calcite equilibria to some of the ground-water analyses may be
useful here. Approximations of this kind often are of
value in water-analysis interpretation.
Analysis 1 in table 15 representsa large limestone
spring, and as might be expected the water evidently is
not far from calcite equilibrium. The calculated S.I.
value for this solution is -0.21. The water representedby
analysis 9 in table 16 also came from a limestone. Although this solution is less dominated by calcium and
bicarbonate, it also is near calcite equilibrium, with an
S.I. of +0.20.
Springsthat form depositsof travertine are common
in many areas of the United States (Feth and Barnes,
1979). Analysis 3, table 11, representsthe water from
such a spring which issuesfrom Redwall Limestone in
the Grand Canyon region of Arizona. Such waters are

mg/L

74

meq/L

.. . .. . .. . . . .

and Jones (1952); 8, U.S. Geological

7
May 19, 1952
mg/L

1948 and 1949

meq/L

4.9

mg/L

meq/L

__________________
17
__................

Survey Water-Supply
9
June 6, 1954

mg/L

meq/L

.02

.._._...____

.oo
99
28
4.1
287
120
6
2.8

4.94
2.30

277
64

13.82
5.26

.18

53

2.29

85
113
605
.2
35

1.39
2.35
17.06
.Ol
.56

4.70
2.50
.17
.05

401
..... 1,260
..................
.......................................................................
362
... 954
127
.................. 884
..................
651
...... 2,340
..................

....................................................................

25
3.9
4.5
1.4
90
12
2.2
.l
1.9

100
99
78
5
172
6.7
5

1.248
,321
,196
,036 >
1.475
,250
,062
,005
.031

394
93

19.66
7.65

333

14.48

___
.................. 2,610
.............. .........................................
..... 1,370
..................
_____
1,240
..................
...... 3,540
...
.......................................................
.............. .........................................

10
Dec. 8, 1954
mg/L

meq/L

15

24

.04

Paper 1163 (p. 360)]

4.39
.60
.83
.07
5.24
.I4
.37
.02
.07

88
7.3
19
2.8

323
322
250
0
543
7.5
2

1.0

.Ol
96
19
18
1.5
133
208
25
.4
.4

4.790
1.562
.I83
.038
2.180
4.330
,705
,021
,006

.................. 2449
.................. 468
... 318
.................. 209
... 690

..................

..................
..................

_.__._____________
..................

7.8
3

..................
..................

5. Rattlesnake Spring, sec. 25, T. 24 S., R 23 E., Eddy County, N. Mex. Flow, 2,500 gpm Water-bearing formatron, alluvrum, probably fed by the Caprtan
Limestone.
6 frrrgation well, NE1/4 set 35, T. I S., R. 6 E., Maricopa County, Ariz. Temperature, 25.6OC. Water-bearing formation, alluvium
7. Cumberland River at Smithland, Ky. Discharge, 17,100 cfs.
8. Pecos Rover near Artesia, N. Mex. Discharge-weighted average, 1949 water year; mean discharge, 298 cfs.
9 Crty well at Bushton, Race County, Kans. Depth, 99 ft. Water-bearing formation, Dakota Sandstone.
10. Industrial well, Wdlimansett, Mass, Depth, 120 ft; temperature, 12.2V. Water-bearing formatron, Portland Arkose.

Significance

of Properties

and Constituents

Reported

in Water

Analyses

95

aswell asby chemical solution or precipitation of minerals.The formation of calcium chloride brine was ascribed
by Valyashko and Vlasova (1965) to a combination of
concentration and ion-exchange processes.
The process of cation exchange in ground-water
bodies has been observedextensively since early studies
by Renick (1924) called attention to natural softening of
ground water by cation exchangein sedimentsunderlying
the northern Great Plains area of the United States.
Ground water that has exchangedcalcium for sodium is
common in many areas(seeanalyses 1 and 2, table 17).
The reverseeffect, exchangeof sodium for calcium, also
can be expectedbut is lesswell documented. In localities
near the ocean where seawater has entered freshwater
aquifers,the advancingsaltwater ,front commonly carries
higher proportions of calcium to sodium than are characteristic of seawater (Poland and others, 1959, p. 193),
owing to releaseof calcium from exchange positions by
the sodium brought in by the advancing saltwater.
In irrigated areas,the exchangeof calcium for sodium
in soil moisture may proceed forward or in reverse at
different times and in any specific spot may fluctuate
extensively.The residualsolutesfrom the irrigation water
that was used by the crop or evaporated from the soil,
however, will partly reappear in drainage water. The
water of the Salt River, which is extensively used for
irrigation in the Phoenix, Ariz., area,is shown by analysis
5, table 17, to contain nearly twice as many milliequivalents per liter of Na as of Ca2+Mg2. The ion
ratio in the effluent, principally drainage of residual
irrigation water, is representedby analysis 6, table 17; it
has a slightly higher proportion of the divalent ions, and

substantially supersaturatedat normal atmospheric pressure. Reaching the calcium activity in this analysis by
solution of calcite would require a partial pressure of
CO2 near 0.4 atmosphere (fig. 18). The S.I. calculated
for this analysis is +0.13. Other travertine-depositing
springs are representedby analysesgiven by White and
others (1963, p. F54).
Analysis 9 in table 15 representswater from sandstone with calcareouscement that hasgiven the solution
a character typical of saturation with calcite. The S.I. for
this solution is +0.28.
Analysis 3 in table 15 is for a spring that issuesfrom
gypsum. From the expressions
[Ca] [so~-+]=llp~~25
and
[CaSo41

=1o2

31

[Ca] [SO?-I
(equilibrium constantsfor 25C from Sillen and Martell,
1964)the saturationindex for this water can be calculated,
by the procedure outlined under the heading Solubility
Product. The S.I. value obtained is -0.025, indicating a
close approach to equilibrium. The effect of temperature
departure from 25OCwas not considered in this calculation but is lessimportant for gypsum that it would be for
calcite equilibria. Evaluations for gypsum saturation pose
few measurementor sample-stability problems, but calculations are tedious if done by hand.
Equilibrium with respectto gypsum is obviously to
be expected in ground water from a gypsiferous aquifer.
The water represented by analysis 3, table 15, is predominantly a solution of calcium and sulfate. As the
amount of other solutesincreases,the solubility of gypsum
will tend to increaseowing to greater ionic strength and
smaller activity coefficients. In a solution containing
2,500 mg/L of chloride and about 1,500 mg/L of sodium,
the equilibrium concentration of calcium would be near
700 mg/L. Thesecalculations presupposenearly equivalent concentrations of calcium and sulfate. Frequently,
this doesnot occur in natural water. When the solution is
in equilibrium with gypsum, however, any increase in
sulfateactivity would be matchedby a decreasein calcium
as CaS04 precipitates (see also fig. 21).
Natural brines in which the predominant dissolved
ions are calcium and chloride are fairly common. An
example is representedby analysis 4 in table 15. There is
no general agreement as to how such a composition is
reached.Water that has beentrapped underground for a
long time could be altered from its original composition
by selectivepermeability of strata for different solutes,by
the bacterial reduction of sulfate and other dissolved
ions, and by adsorption or desorption of dissolved ions,
96

Study

and

Interpretation

of the

Chemical

Characteristics

this suggests that the exchange reactions are occurring,


even though a good deal of the calcium present in the

initial water probably is precipitated as carbonate or,


possibly, sulfate in the irrigated area. The effect is more
strongly shown by analysis 6, table 15, which represents
water from a well in the area irrigated with Salt River
water, where rechargeis composedof irrigation drainage
water. The milliequivalents per liter concentration of
sodium in the water is far lower than that of calcium.
Water that is strongly influenced by irrigation return
flow may approach simultaneous equilibrium with both
calcite and gypsum. Water of the Gila River at Gillespie
Dam near Gila Bend, Ariz., and the Rio Grande at Fort
Quitman, Tex., shows these properties at times (Hem,
1966). Whether analysis 3, table 15, depicts equilibrium
with both solids cannot be determined becausethe pH is
not known.
Magnesium

Magnesium is an alkaline-earth metal and has only


one oxidation state of significance in water chemistry,
Mg2. It is a common element and is essential in plant
and animal nutrition.
of

Natural

Water

In some aspects of water chemistry, calcium and


magnesium may be considered as having similar effects,
asin their contributions to the property of hardness.The
geochemical behavior of magnesium, however, is substantially different from that of calcium. Magnesium ions
are smaller than sodium or calcium ions and can be
accommodated in the spaceat the center of six octahedrally coordinatedwater molecules,an arrangementsimilar
to that described for aluminum. The hydration shell of
the magnesium ion is not as strongly held as that of
aluminum ions, but the effect of hydration is much
greaterfor magnesiumthan for the larger ions of calcium
and sodium. The tendency for precipitated crystalline
magnesiumcompounds to contain water or hydroxide is
probably related to this hydration tendency.
Sources

of Magnesium

In igneous rock, magnesium is typically a major


constituent of the dark-colored ferromagnesianminerals.
Specifically, these include olivine, the pyroxenes, the
amphiboles, and the dark-colored micas, along with
variouslesscommon species.In altered rocks, magnesian
mineral species such as chlorite and serpentine occur.
Sedimentary forms of magnesium include carbonates
such as magnesite and hydromagnesite, the hydroxide
brucite, and mixtures of magnesium with calcium carbonate.Dolomite hasa definite crystal structure in which
calcium and magnesium ions are present in equal
amounts.
The alteration of magnesianolivine (forsterite) to
serpentinite can be written

This is somewhat analogous to the reactions shown


previously for weathering of feldspar and produces a
solid alteration product, serpentinite. The reaction, like
that for the alteration of other silicates, is not reversible
and cannot be treatedasa chemical equilibrium. Released
products, however, can be expected to participate in
additional reactions.
Form

of Dissolved

Magnesium

The magnesium ion, Mg, will normally be the


predominant form of magnesium in solution in natural
water. Data given by Sillen and Martell (1964, p. 41-42)
show that the complex MgOH will not be significant
below about pH 10. The ion pair MgSO.,(aq) has about
the samestability asthe speciesCaS04(aq), and magnesium complexes with carbonate or bicarbonate have
approximately the samestability asthe similar speciesof
calcium. The sulfateion pair and the bicarbonatecomplex
Significance

will be significant if the solution contains more than


1,000 mg/L of sulfate or bicarbonate.
Chemical

Controls

of Magnesium

Concentration

Carbonateequilibria involving magnesiumare more


complicatedthat thosegoverningcalcium activities. There
are severaldifferent forms of magnesiumcarbonatesand
hydroxycarbonatesand they may not dissolve reversibly.
Magnesite, MgC03, from solubility products given by
Sillen and Martell (1964, p. 136,137), seemsto be about
twice assoluble ascalcite. However, the hydrated species
nesquehonite, MgC03*3HzO,
and lansfordite,
MgC03*5Hz0, are considerably more soluble than magnesite.The basic carbonatehydromagnesite, Mg,(COz)s
(OH)z.3Hz0, may be the least soluble species under
some conditions. Magnesite apparently is not usually
precipitated directly from solution, and, as noted by
Hostetler(1964), a considerabledegreeof supersaturation
with respectto all magnesiumcarbonate speciesmay be
required before precipitation can occur. In any event,
such species are rarely significant factors in limiting
magnesiumcontent of water.
Magnesium occurs in significant amounts in most
limestones. The dissolution of this material obviously
brings magnesium into solution, but the process is not
readily reversible-that is, the precipitate that forms
from a solution that hasattacked a magnesianlimestone
may be nearly pure calcite. Magnesium concentration
would tend to increasealong the flow path of a ground
water undergoing such processes,until a rather high
[Mg]:[Ca] ratio is reached.
Conditions under which direct precipitation of
dolomite from solution occurs are not commonly found
in natural-water (nonmarine) environments. Dolomite
precipitates have been found in saline lakes in various
places. Hostetler (1964) cited reports of its occurrence
from South Australia, Austria, Utah, and the U.S.S.R.
but noted that the concentrationsof magnesium,calcium,
and carbonateions in water associatedwith the dolomite
seemedto be abovetheoretical saturation limits in many
instances.
Recent workers seem to be in fairly general agreement that the solubility product for dolomite is near
1O-17oat 25 C. Langmuir (1971) accepted this value
and calculated valuesof lo-l6 56at O, lo-l6 71at lo, and
10-6.8gat 20C. Earlier, severalinvestigators (Hsu, 1963;
Barnesand Back, 1964a;Holland and others, 1964) used
analysesof ground water in association with dolomite to
calculatean ion-activity product near lo-l6 5.The general
characteristics of ground water from dolomitic terrane
suggestthat saturation with respectto calcite also occurs
in many such waters. Yanateva (1954) reported that
dolomite was somewhat more soluble than calcite under
a partial pressureof CO2 approaching that of ordinary
air.
of Properties

and Constituents

Reported

in Water

Analyses

97

by clay minerals and other surfaces having exchange


sites.
Some magnesium silicate minerals can be synthesizedreadily in the laboratory at 25C. Siffert (1962) for
example, prepared sepiolite (Mgs!%Ort*4HaO) by precipitation from solutions of Si(CrH)d and MgClz at a pH
of 8.73 or higher. This mineral or a related speciesmay
have a significant effect on magnesium concentration in
weathering solutions acting on magnesium-rich igneous
rock.
Eaton and others(1968) proposedthat the solubility
of magnesium in irrigation drainage water might be
limited by precipitation of a magnesium silicate of unspecified composition in the soil. A magnesium sink is
neededto explain the lossof magnesiumthat is commonly
observed in such waters.
Table7, which showsresidencetimes of the elements
in the ocean,points up another factor of someimportance:
Magnesium has a very long residence time compared

The attainment of a relatively stable value for the


activity product is possiblein a systemin which calcite is
precipitated while dolomite dissolves. Long residence
times of water in such systemswould produce [Mg]:[Ca]
ratios above 1.0 and a high pH, with both tending to
increasealong the flow path. This mechanism is treated
quantitatively in the mass-balancemodel that Plummer
and Back (1980) applied to the Floridan and the Madison
Limestoneaquifers.Carbon dioxide could be replenished
within thesesystemsthrough sulfate reduction and, theoretically at least, very high magnesium concentrations
might ultimately be reached. Magnesium hydroxide
(brucite) occurs naturally, and the solubility product at
25C is between IO- and 10-r (Sillen and Martell,
1964,p. 41-42). Precipitation of this solid implies a high
pH and a solution impoverished in dissolved carbon
dioxide species.
The cation-exchangebehavior of magnesiumis similar to that of calcium. Both ions are strongly adsorbed

Table 16. Analyses of waters in which

[Analysesby

U.S. Geological Survey except as indicated. Date under sample number IS date of collection.

Constituent

Feb. 27,
mg/L

Silica (SiOz) ............................


Aluminum (Al) ........................
Iron (Fe) ..................................
Calcium (Ca) ............................
Magnesium (Mg) ......................
Sodium

(Na)

............................

1952

April

meq/L

mg/L

1.996
1.809
,017
.03l
3.491
,102
,056
,000
,077

1.4
94
40
17
2.2
0
471
49
9.0
.8
2.4

4.69
3.29
.74
.06
____._.............

7.72
1.02
.25
.04
.04

mg/L

Oct.

meq/L

31

_...................

20
42

1.oo
3.45

19

.83

>
0

279
22
7
.2
2.5

mg/L
175

34
242
184
18

18, 1957

meq/L
.. .... . .... . . . ..

1.70
19.90
8.00
.46 >

4.57
.46
.20
.Ol
.04

1,300
6.6
265
1.0
.2

21.31
.l4
7.47
.05
.oo

__.................. 281

_.___.......__._____
1,580

._________..........

...................

...................

.............
....................
....................

...............
....................

....................

6.7
5

Sept. 29, 1948

meq/L

466
527
400
13
764

Includes 18 mg/L of boron (B).


*Contains 0.01 mg/L of manganese (Mn) and 0 9 mg/L of lithium
3Denstty, 1,345 g/mL.

98

11, 1952

18

8.4
1.4
.24
40
22
.4
1.2
0
213
4.9
2.0
.O
4.8

Potassium (K) ..........................


Carbonate (COa) ......................
Bicarbonate (HCO$ ................
Sulfate (SOI) ............................
Chloride (Cl) ............................
Fluoride (F) ..............................
Nitrate (NOs) ..........................
Dissolved solids:
Calculated ............................
190
Residue on evaporation ________180
Hardness as CaC03 ..................
190
Noncarbonate ......................
16
Specific conductance
326
(micromhos at 25T).
pH ............................................
7.4
Color ........................................
5

1.
2.
3.
4.

Sources of data: I,3 -7, and 9, U.S. Geological Survey,

222
0
458
8.2

...................
...................

1,080
14
2.500
6.5

............

(Li).

Spring 2% mi northwest of Jefferson Cuy, Tenn. Flow, 5,000 gpm; temperature, 14 4C. From the Knox Dolomne.
Well number 5, City of Sidney, Ohio. Depth, 231 ft; temperature, 11.7% Water-bearing formatron, Niagara Group.
Spring in Buell Park, Navajo lndran Reservation, Ariz. Flow, 18 gpm; temperature 12.2C. Water-bearmg formatton, olivine tuff-breccta.
Mam spring at Siegler Hot Springs, NEl/4 sec. 24, T. 12 N., R. 8 W., Lake County, Calif. Water issues from contact wrth serpentine and sedrmentary rocks.
Temperature, 52.SC.
Study

and

interpretation

of the

Chemical

Characteristics

of

Natural

Water

with calcium. Organisms that require calcium for shell


or skeletal parts representan important factor in calcium
precipitation. Such demands on magnesium are much
smaller.

Occurrence

of Magnesium

should dissolve equal molar amounts of calcium and


magnesium. Analyses 1 and 2 in table 16 have negative
valuesfor S.I. for calcite and for dolomite. Analysis 1 has
nearly equal meq/L valuesfor calcium and magnesium.
The greater proportion of calcium in analysis 2 may
reflect the fact that the rocks of the Niagara Group, from
which the water comes, include limestone as well as
dolomite.
Water represented by analysis 10, table 16, has
substantial supersaturation with respect to calcite
(S.I.=+O.40) and a calculated S.I. for dolomite of +0.99.
Dissolved magnesiumexceedscalcium in this water, and
this canbe explainedby calcite precipitation after reaching
supersaturation. As noted earlier, dolomite does not
precipitate readily under theseconditions. Analysis 5 in
table 16illustrates this behavior even more strongly. This
water is from a pool in Carlsbad Caverns, N. Mex., in
which calcite is precipitating. The dolomite S.I. in this
solution is +1.58.

in Water

Analyses in table 16 can be used to demonstrate


someof the effectsof dolomite dissolution on magnesium
content and other properties of natural water. As noted
earlier, the calculation of S.I. values from these data is
subject to an unknown but probably substantial uncertainty because the reported pH values are laboratory
determinations. Analyses 1, 2, and 10 representground
waters from dolomitic terranes and analysis 9 is for
ground water from limestone containing dolomite.
Before it reachessaturation with respect to either
dolomite or calcite, a water passing through dolomite
magnesium

is a major

constituent

unpublished data; 2, 8, and 10, Scott and Barker (1962, p. 19, 87, 113)]

Feb. 7, 1939

May 21, 1952

May 27, 1952

5
Apr.
mg/L
11

20, 1945
meq/L

mg/L

meq/L

........................................................

.04 ........................................................
16
30
130
6.49
71
5.84
51,500
4,236.39
59,000
2,566.60
7.8
.34
71.85
{ 2,810
21
....................................
.70
320
1,860
5.25
30.49
21
.44 299,000
6,225.18
8
22,600
.23
637.55
1.0
....................................
.05
19
....................................
.31
333
321
332
34
570

....... -436,000
..................
..................
.................. 510,000
........................................................
........................................................
(q
..................

8.3 .......................................................
..........................................................................

mg/L

mg/L

meq/L

4.2
.I3
18

8.5
2.5
2.4
0
90
9.5
1.5
.O
.8
92
106
80
6
165
7.0
35

..................
............ .....
,898
.699
.109
,061
..................
1.475
,198
,042
.ooo
.013

..
..................
..................
..................
..................

meq/L

Mar.

9
11, 1952
meq/L

mg/L

12
1.0
.03
23
12
1.7
1.6
0
127
1.6
2.0
.O
.9

13
..................
..................
........................................................
..................
.Ol ..................
1.148
6.99
140
.987
43
3.54
,074
21
.91
,041
{
>
..................
0
241
3.95
2.08 1
.033
6.31
303
,056
1.07
38
,000
.8
.04
,014
.07
4.1

119

..................

108
107
3
197

..................
..................
..................
..................

7.6
5

682
701
526
329
997

10
Oct. 29, 1954
mg/L
18

meq/L
...............

.2 ...............
.39 ...............
35
1.747
2.714
33
28
1.218
1.3
.033
0
.ooo
241
3.950
88
1.832
,028
1.0
.9
,047
1.2
,019

.... 326
.................. 329
.................. 224
..................
27
.................. 51 1

...............
...............
...............
...............
...............

..................
7.4 ..................
8.2 ...............
.........................................................................................

5.
6.
7.
8.

Green Lake in Carlsbad Caverns, N. Mex. Pool of ground-water seepage.


Test well one, SW]/4 sec. 24, T. 25 S., R. 26 E., Eddy County, N. Mex. Water from 142-195 ft.
Wisconsin River at Muscoda, Grant County, Wis. Flows through area of magnesian limestone.
Spring, SW]/4 sec. 26, T. 16 S., R. 7 E., Calhoun County, Ala. Supplies city of Anniston. Flow, 46 cfs; temperature, 17.goC. Water-bearing formation,
quartzite.
9. Oasis flowing well, SW114 sec. 15, T. 11 S., R. 25 E., Chaves County, N. Mex. Depth, 843 ft; flow, more than 9,000 gpm when drilled. Water-bearing
formation, San Andreas Limestone (hmestone and dolomitic limestone, with minor amounts of sandstone, gypsum, and anhydrlte). Temperature, 20.6OC.
10. Drilled well, NW1/4 sec. 6, T. 6 N., R. 21 E., Milwaukee County, Wts. Depth, 500 ft; temperature, 10.OoC. Water-bearing formation, Niagara Dolomite,
Significance

of Properties

and Constituents

Reported

in Water

Analyses

99

concentrationsin water, in the way that carbonateprecipitation controls calcium concentrations. Sodium is retained by adsorption on mineral surfaces,especially by
minerals having high cation-exchangecapacities such as
clays. However, the interaction between surfacesitesand
sodium, and with monovalent ions generally, is much
weaker than the interactions with divalent ions. Cationexchangeprocessesin freshwater systemstend to extract
divalent ions from solution and to replace them with
monovalent ions.

A similar effect may be noted by comparing analyses


1 and 2 in table 11. These representbefore and after
composition of a water that lost calcium from solution
by formation of a calcium carbonate precipitate inside
the sample bottle during storage.
Analysis 9 in table 16 represents water from the
Roswell artesian system in southeastern New Mexico.
This water is in contact with limestone and dolomite and
appearsto be near saturation with respectto both calcite
and dolomite; computed values for S.I. are +0.20 and
+O.18, respectively.
Analysis 7 in table 16 represents water from the
Wisconsin River. The rather high proportion of magnesium to calcium in streamsof that region was pointed
out many years ago by Clarke (1924a).
The chemical reaction written earlier in this discussion for the conversion of olivine to serpentinite indicates
that the processconsumes much H. This could also be
consideredequivalent to production of OH-. Barnesand
others(1967) called attention to ground waters of unusual
compositionthat they observedin association with fresh
or partly serpentinizedperidotite or dunite. An example
is given as analysis 5, table 18. This water had a pH of
11.78 and carbonate specieswere below detection. The
magnesiumconcentration is very small owing to precipitation of brucite (Mg(OH)z). At the calcium concentration
reported in the analysis, only a few tenths of a milligram
of cos2- per liter could be present at equilibrium with
CaC03. Barnes,ONeil, and Trescases(1978) have cited
other examplesof theseultrabasic waters, from Oman,
New Caledonia, and Yugoslavia.The hydroxide concentration in analysis5, table 18,was calculated from the pH

According to Clarkes (1924a, p. 6) estimate, about


60 percent of the body of igneous rock in the Earths
outer crust consists of feldspar minerals. The feldspars
are tectosilicates, with some aluminum substituted for
silica and with other cations making up for the positivechargedeficiency that results.The common feldsparsare
orthoclase and microcline, which have the formula
KAlSisOs, and the plagioclaseseriesranging in composition from albite, NaAlSisOs, to anorthite, CaAl&isOs.
Some sodium may be present,substituting for potassium
in orthoclase and microcline.
Potassium feldspar is resistant to chemical attack.
However, species containing sodium and calcium are
somewhat more susceptibleto w,eathering;they yield the
metal cation and silica to solution and commonly form a
clay mineral with the aluminum and part of the original
silica. Besidesthe kaolinite shown as a product in earlier
discussionsof this reaction in this book, other clays such
as illite or montmorillonite may be formed. Feth and

value, which

others (1964),

was measured

in the field

at the time

of

samplecollection. Someof the calcium is probably present


as the CaOH complex, and the silica and aluminum
would be present in anionic form at this pH. With these
factors taken into account, the cation-anion balance is
satisfactory.
Sodium

Sodium is the most abundant member of the alkalimetal group of the periodic table. The other naturally
occurring membersof this group are lithium, potassium,
rubidium, and cesium.In igneousrocks, sodium is slightly
more abundantthan potassium,but in sediments,sodium
is much less abundant. The amounts of sodium held in
evaporite sediments and in solution in the ocean are an
important part of the total. All the alkali metals occur in
the l+ oxidation state and do not participate in redox
processes.Sodium ions have a radius somewhat greater
than 1 angstrom and are not strongly hydrated.
When sodium has been brought into solution, it
tends to remain in that status. There are no important
precipitation reactions that can maintain low sodium
100

Study

and Interpretation

of the Chemical

Characteristics

Sources

of Sodium

in studies

of water

in the feldspar-rich

granitic terrane of the Sierra Nevada of California and


Nevada, found that, in general,calcium and sodium ions
were the most important cations in stream and spring
water and that these tended to reflect abundance of the
ions in the type of rock and the rate at which the minerals
were attacked.
In resistate sediments, sodium may be present in
unaltered mineral grains, asan impurity in the cementing
material, or as crystals of readily soluble sodium salts
deposited with the sediments or left in them by saline
water that entered them at some later time. The soluble
salts go into solution readily and are rather quickly
removed from coarse-grained sediments after environmental changes,such as an uplift of land surface or a
declineof sealevel, imposea freshwaterleaching regimen.
During the early stagesof the leaching process,the water
leaving the formation may have high concentrations of
sodium in solution. The last traces of marine salt or
connatewater may persistfor long periods where circulation of water is impaired.
In hydrolyzate sediments, the particles normally
are very small, and the circulation of water through the
of Natural

Water

material is impaired. Thus, the water trapped in the


sediment when it was laid down may be retained with its
solute load for long periods. The hydrolyzates include a
large proportion of clay minerals having large cationexchange capacities. Where sands and clays are interbedded,water and sodium may be retained longer in the
less permeable strata during leaching and flushing by
freshwater circulation. When such interbedded sections
are penetrated by wells, water will be drawn mainly
from more permeable sections at first. Long-continued
withdrawals and water-table declinescan be expected to
alter water-circulation patterns, and saline solutions can
be induced to move from the clay and shalelayers. When
this occurs on a large scale, a substantial increase in
sodium concentration of the pumped water will occur.
Kister and Hardt (1966) observedthis effect in irrigation
wells of the SantaCruz basin in Arizona, where extensive
declines in water levels had occurred.
Human activities can have a significant influence
on the concentrations of sodium in surface water and
ground water. The use of salt for deicing highways in
winter and the disposal of brine pumped or flowing from
oil wells, for example,havehad direct, noticeableregional
effects. Somewhat less directly, the reuse of water for
irrigation commonly leavesa residual that is much higher
in sodium concentration than was the original water.
Pumping of ground water, which alters hydraulic gradients,can induce lateral movement of seawaterinto freshwater coastal aquifers.
Ion-exchangeand membraneeffectsassociatedwith
claysweredescribedunder the topic Membrane Effects.
Some of these factors may influence sodium concentrations. Hanshaw (1964), for example, showed that when
compacted, clays may preferentially adsorb sodium, but
when dispersedin water, they may preferentially adsorb
calcium.
Dissolved

Species

The sodium of dilute waters in which dissolvedsolids concentrations are below 1,000 mg/L is generally
in the form of the Na ion. In more concentratedsolutions,
however, a variety of complex ions and ion pairs is
possible. Speciesfor which stability constants are given
in Sillen and Martell (1964, p. 136,235) include NaCd:;,
NaHCOa(aq), and NaS04. Theseand other ion pairs or
complexes involving sodium are substantially lessstable
than the ones involving divalent cations such as calcium
or magnesium.
Solubility

Controls

Becauseof the high sodium concentrations that can


be reachedbefore any precipitate is formed, the sodium
concentrations in natural water can have a very wide
range, from less than I mg/L in rainwater and dilute
stream runoff in areasof high rainfall to very high levels
Significance

in brines associatedwith evaporite deposits and in brines


of closedbasins,where more than 100,000 mg/L may be
present.
Sodium bicarbonate is one of the lesssoluble of the
common sodium salts. At ordinary room temperature,
however, a pure solution of this salt could contain as
much asabout 15,000 mg/L of sodium. In natural water,
the conditions required for precipitation of pure sodium
bicarbonate are unlikely to be attained, although water
in some closed basins may attain high concentrations of
carbonate and bicarbonate and leave a residue of solid
forms of sodium carbonate. These are somewhat more
soluble than the bicarbonate.
Sodium carbonate residual brines occur in some
closed basins in California, Oregon, and Washington
and elsewhere.Garrels and MacKenzie (1967) described
a sequenceof concentration and precipitation of solids
from water that originally obtained its solute contents
from weathering of silicate minerals in igneous rock.
These processesmay lead to solutions having a pH of
more than 10.
A higher solubility limit on sodium concentration is
exerted by the separation of solid sodium chloride, or
halite. When saturated with respect to halite, a solution
could have as much as 150,000 mg/L of sodium and
about 230,000 mg/L of chloride, but concentrations this
high are seldom reachedin natural environments. Inclusions of brine are present in some halite formations. It
can be shown, theoretically at least, that such inclusions
may be able to migrate through the salt if a temperature
gradient exists, by dissolving salt at the warmer end of
the inclusion and precipitating it at the cooler end.
Sodium sulfate solubility is strongly influenced by
temperature. The solid precipitated may contain various
amounts of water, ranging from mirabilite or Glaubers
salt with the formula NaaS04*lOHaO, through the
heptahydate with seven molecules of water and the
anhydrous form. Closed-basin lakes in cool climates
may precipitate mirabilite during cool weather, which
may be redissolved at higher temperatures. Mitten and
others (1968) discribed such effects in eastern Stump
Lake, N. Dak. Sodium concentrations in the lake during
a 5-year period of intermittent sampling generally were
between 20,000 and 30,000 ppm. An apparent decrease
of about 25 percent in sodium concentration and a
corresponding lossof sulfate was reported over a l-week
period when the water temperature decreasedfrom 11
to 3C (Mitten and others, 1968, p. 26). Somewhat
similar deposition of mirabilite has been observed in
Great Salt Lake, Utah (Eardley, 1938).
Occurrence

of Sodium

in Water

In table 17 and some other places in this book are


analyses that illustrate the various features of sodium
behavior in natural aqueoussystems.
of Properties

and Constituents

Reported

in Water

Analyses

101

analysis 4, table 17. This analysis representsa stream in


western South Dakota where the country rock is fine
grained and contains soluble material. In the processof
weathering, the exposed surfaces of these rocks may
develop noticeable efflorescences,of salts by evaporation
of water between rainy periods. Aqueous precipitation
heavy enough to wash away thesedeposits occurs occasionally, and at the same time the layer of leached
sedimentat the surfacealso may be stripped away. Colby
and others (1953) have discussedtheseeffects in greater
detail.
Wells that penetrate clay and shale to reach lower,
more permeable beds may at times obtain water high in
dissolved solids directly from the shale layers, as seems
indicated in analysis 9, table 17. Analysis 8, table 14,

Analysis 2 in table 12representswater from rhyolitic


volcanic rock and indicates the predominance of sodium
over other cations that ought to be observedin water that
has attacked albite. Waters that are low in dissolved
solids, and in which calcium and magnesium have been
depletedand sodium increasedthrough cation exchange,
are common in sediments of the Atlantic Coastal Plain
and the Mississippi embayment. Analysis 2 in table 14 is
for a dilute water from the Wilcox Formation showing
this effect. A more strongly influenced example is analysis
1 in table 17, which representsa well in Atlantic Coastal
Plain sediments. Analysis 2 in table 17 also shows the
influence of cation exchange.
Some effectson surface-streamchemistry of solutes
retained in fine-grained sediment are demonstrated by

Table 17. Analyses


[Analyses by U.S. Geological Survey. Date below sample number is date of collection.
214); 9, Griggs and Hendrickson (1951, p. Ill)]

2
Oct. 3, 1949

Constituent

June 3, 1952
mg/L

Silica (SiOa) .......I ....................


Iron (Fe) ..................................
Calcium (Ca) ............................
Magnesium (Mg) ......................

meq/L

.12
.17

..........................

7.90

>

.15
3.0
7.4
857
2.4

Carbonate
(CO3)...................... 30
Bicarbonate (HCOa) ................
412
Sulfate (Sod) ............................
3.5
Chloride (Cl) ............................
9.5
Fluoride (F) ..............................
1.7
Nitrate (NOa) ..........................
.6
Boron (B) .....................................................
Dissolved solids:
Calculated ............................
457
Residue on evaporation ._______ 452

Hardness
asCaC03..................
Noncarbonate ......................
Specific conductance
(micromhos at 25C).
pH ............................................
Color ........................................
Density,
Density,

1.019 g/mL
1.21 g/mL.

15
0
718

38
.
0
_._......_________..
2,960

8.7
1

8.3

mg/L

22
.
__......______......
.oo
.15
49
61
18
37.28
168
.06
1.90
0
34.09
202
.03
44
2.00
246
.I1
.l
2.2
.oo

._.______... 2,060

__
.........................
....................
....................
....................

....................
........................

Study

7.7
2

meq/L

2.45
I .48
7.29

4
July 2-3,

mg/L
8.2
.04
40
50
699
16

1949

meq/L

2.00
4.11
30.40
.41
.87
7.47
27.48
.48
.05
.03

.. . .._..______._..2,400
.
2,410
.
306
0
3,140
8.2

at 20C.

1. Well at the Raleigh-Durham


Airport, Wake County, N.C. Depth, I84 ft. Water-bearing formation,
2. Well, SE114 NE114 sec. 2, T. 22 N., R. 59 E., Richland County, Mont. Depth, 500 ft. Water-bearing
3. Irrigation well, SE1/4 sec. 3, T. 1 N., R. 5 E., Maricopa County, Ariz. Depth, 500 ft.; temperature,

102

649
651
196
31
1,200

in which

and 8, U.S. Geological Survey, unpubhshed data; 2,

3
May 13, 1952

meq/L

16

Sodium(Na) ............................t ra2


Potassium (K)

mg/L

22
.20
2.5
2.1

Sources of data: 1,3,7,

of water

and interpretation

of the Chemical

Characteristics

of Natural

Water

Coastal Plain sedimentary rocks.


formation, Fort Union Formation (sandstone and shale)
20.5C. Water-bearing formation, valley fill.

showsthe composition of water obtained from the Dakota


Sandstonein southeastern North Dakota. The water is
high in sodium and in sulfate. Solute contents of water in
the formation in this region may be derived from overlying
and underlying finer grained materials (Swenson, 1968).
Water associatedwith evaporiteformations generally
hasa very high sodium concentration. Analysis 2 in table
18 representsa sodium carbonate brine from Wyoming.
Sodium carbonateevaporitedepositsoccuring in southern
Wyoming are mined as a source of soda ash.
Water affected by solution of halite is represented
by analyses 7 and 8 in table 17. Analysis 8 displays
near-saturation with respect to sodium chloride. The
source of the salinity in the Salt Banks spring is not
evident at the surface. Inflows of this type have a signifi-

sodium

is a major

cant influence on the water of the Salt River, which is


impounded for usein irrigation of the area near Phoenix,
Ariz. The composition of this supply is represented by
analysis 5 in table 17.
Anthropogenic effects on sodium concentrations
are demonstratedby high sodium concentration of residual drainage from the Salt River Valley irrigated area
(analysis 6, table 17). Somewhat higher averageconcentrations occurred at this sampling point during the 10
yearsfollowing 1952.Most of the increasewas in sodium,
chloride, and sulfate. The water at low flow is near
saturation with respect to calcite and gypsum (Hem,
1966). Analysis 3, table 17, represents a well in the
irrigated area upstream where recharge is virtually all
brought about by applied irrigation water. The analysis

constituent

Torrey and Kohout (1956, p. 44); 4, U.S. Geological Survey Water-Supply Paper 1162 (p. 457); 5 and 6, U.S. Geological Survey Water-Supply Paper 1253 (p. 205,

5
1951 -52
mg/L

18
.Ol
48
14
150
5.8
0
153
50
233
.4
2.4
.15
597
611
178
52
1,090

6
1951 -52

meq/L

mg/L

....................
....................
2.40
1.15
6.52
.15

31

2.51
1.04
6.57
.02
.04

.Ol
353
149
1,220
9.8
0
355
1,000
1,980
1.9
24
2.4
4,940
1,490
1,200
7,620

meq/L

7
Dec. 8, 1951
mg/L

meq/L

a
Jan. 31, 1938
mg/L

meq/L

9
Oct. 15, 1946
mg/L

meq/L

....................
46
................................................................
13
....................................................................................................................................................
17.61
505
25.20
722
36.03
30
12.25
291
23.94
2,490
204.83
31
53.05
10,100
439.35
121,000
5,263.50
279
.25
170
4.35
3,700
94.61 >
0
63
2.10
0
5.82
1,520
24.91
40
.66
445
20.82
899
18.72
11,700
243.59
303
55.84
16,200
457.00
189,000
5,331.69
80
.lO
1.2
.39
.. . . . . . .. . . .. . . . .. . . . . .. . . . .. . . .. . . .. . . . .. .. . .. .. .
17
17
29,000

.._._329,000

2,460
1,210
_____41,500

____________._______
225,000

1.50
2.55
12.13

7.29
6.31
2.26
.06
.27

973
202
0
____________________
1,510

..__..___._...___.__

7.1

4. Moreau River at Bixby, S. Dak.; composite of two daily samples.Mean discharge, 1 7 cfs. Drams the Pierre Shale, the Fox Hills Sandstone,and the Hell Creek
Formatlou.
5. Salt Rwer below the Stewart Mountain Dam, Ariz. Discharge-weighted average, 1952 water year; mean discharge, 362 cfs.
6. Gila River at the Gillespie Dam, Ariz. Discharge-weighted average, 1952 water year; mean discharge, 71 I cfs.
7. Spring entering the Salt River at Salt Banks near Chrysotile, Ariz. Temperature, 21 1% Water-bearing formation, quartzite and dlabase.
8. Test well 3, sec. 8, T. 24 S., R. 29, E., Eddy County N. Mex. Depth, 292 ft. Brme from the Salalado Formation and the Rustler Formation.
9 Well in SW1/4 sec. 7, T. I7 N., R. 26 E., San Miguel County, N. Mex. Depth, 50 ft. Water-bearing formation, shale of the Chinle Formation.
Significance

of Properties

and Constituents

Reported

in Water

Analyses

103

is nearly the same as that for the river water (analysis 5,


table 17). The effect of seawater encroachment in a
coastal aquifer near Los Angeles, Calif. is shown by
analysesin table 22.
Many water analysesin the earlier literature report
a computed value for sodium rather than an actually
determined one. The computed value representsthe difference between the sum of the determined anions, in
milliequivalents per liter, and the determined cations,
expressedin the sameunits. Obviously, the computation
cannot be made unlessall the other major ions have been
determined.Potassiumis commonly lumped with sodium
in the computation, and the value reported as sodium
and potassium, as sodium. The principal reason for
omitting sodium from the analysis in former years was
that the determination was difficult, tedious, and expensive by the proceduresthen available. After about 1955,
the flame photometer and related flame spectrochemical
methods made the sodium determination one of the
quickest and easiestin the analytical chemists repertoire,
and, thus, analyseswith computed sodium values are no
longer common. The principal objection to the computation of sodium is that it prevents any really effective
check of the accuracy of the determinations of the major
dissolvedcomponents;therefore,analysesreporting computed sodium values are more likely to contain major
errors.
Potassium

As shown in table 1, potassium is slightly lesscommon than sodium in igneous rock but more abundant in
all the sedimentary rocks. In the oceanthe concentration
of potassium, though substantial, is far less than that of
sodium. The figures point up the very different behavior
of these two alkali metals in natural systems. Sodium
tends to remain in solution rather persistently once it has
beenliberated from silicate-mineral structures.Potassium
is liberated with greater difficulty from silicate minerals
and exhibits a strong tendency to be reincorporated into
solid weathering products, especially certain clay minerals. In most natural water, the concentration of potassium
is much lower than the concentration of sodium.
Another important factor in the hydrochemical behavior of potassium is its involvement in the biosphere,
especiallyin vegetationand soil. Potassiumis an essential
element for both plants and animals. Maintenance of
optimum soil fertility entails providing a supply of available potassium. The element is present in plant material
and is lost from agricultural soil by crop harvesting and
removal as well as by leaching and runoff acting on
organic residues.
Sources

of Potassium

Study

and

Interpretation

of the

Chemical

Characteristics

Control

Mechanisms

for

Potassium

Concentration

Although potassium is an abundant element and its


common saltsare highly soluble, it seldom occurs in high
concentrations in natural water. Chemical mechanisms
that might be expected to bring about such a result,
however, are not of a type that is readily quantified. The
broad generalizations already stated suggestthat potassium concentrations in water are low partly becauseof
the high degreeof stability of potassium-bearingaluminosilicate minerals. Unaltered potassium feldspar grains
occur in many sandstones.
The potassium ion is substantially larger than the
sodium ion, and it would normally be expected to be
adsorbed less strongly than sodium in ion-exchange reactions. Actually, however, potassium is incorporated in
a special way into some clay-mineral structures. In illite,
potassium ions are incorporated in spaces between crystal

layers, and they cannot be removed by further ionexchangereactions.


Potassium ions assimilated by plants become available for re-solution when the plants mature and die, or
when leaves and other parts are shed at the end of the
growing season.In the natural recycling that occurs in
forests and grasslands,this potassium is leached into the
soil by rains during the dormant seasonor made available by the gradual decay of the organic material. Some
leakageof potassium to ground water and runoff during
these processeswould be expected.
There are indications that biological factors may be
important in controlling the availability of potassium for
solution in river water and ground water. Records of
quality of surface waters of the IJnited States show that
at times of relatively high water discharge many streams
in the central part of the country carry potassium concentrations nearly as high as (or higher than) they do at
times of low discharge. This may be the result of soil
leaching by runoff. Similar effects were noted by Steele
(1968b), in Pescadero Creek, Calif., and by Kennedy

The principal potassium minerals of silicate rocks


are the feldsparsorthoclase and microcline (KAlS&OH),
104

the micas, and the feldspathoid leucite (KAlSizOs). The


potassium feldspars are resistant to attack by water.
Presumablythey are altered to silica, clay, and potassium
ions by the same processas other feldspars, only more
slowly.
In sediments,the potassiumIcommonly is presentin
unaltered feldspar or mica particles or in illite or other
clay minerals. Evaporite rocks may locally include beds
of potassium salts and constitute a source for high potassium concentration in brines.
Average content of potassium in living plants, according to data given by Mason (1952, p. 199), is near
0.3 percent. Concentrations in dry plant material and in
ashare substantially greater. Wood asheshave been used
by humans as a potash source for many centuries.

and Malcolm (1977), who made very detailed studies of


of

Natural

Water

discharge and solute contents of the Mattole River in


northern California. Slack (1964) observed increased
potassium in water of pools in the Cacapon River of
West Virginia during autumn, when the water was affected
by the leaching of recently fallen tree leaves.
Some closed-basin lakes in the Sand Hills region of
northern Nebraska contain water having notable concentrations of potassium (Clarke, 1924a) and were considered
possible domestic commercial sources of potash (Hicks,
1921) before the discovery of potassium-rich evaporite
deposits in New Mexico. The reasons for potassium
accumulation in these lake waters have not been fully
established. Hicks (1921) hypothesized that the leaching
of potassium from the ashes of grass destroyed in prairie
fires was a possible source. Such an exotic mechanism is
perhaps unnecessary. The concentration of potassium in
at least some of the near-surface strata in the entire High
Plains region of western Nebraska and Kansas is greater
than the sodium concentration, as indicated by rock
analyses tabulated by Hill and others (1967). Ground
water in the Ogallala and some other formations in the
region may also have relatively high proportions of potassium to sodium (Johnson, 1960). These waters are uniformly low in dissolved solids, as the rock formations do
not contain readily soluble minerals. However, where
water circulation is impaired and the water table intersects
the surface, evaporation could eventually build up potassium concentrations in the water to levels observed in the
Nebraska lakes. Mobilization of potassium by grasses
growing on the highly permeable soil may also play a
significant role.
Occurrence

of

Potassium

in Water

In dilute natural waters in which the sum of sodium


and potassium is less than 10 mg/L, it is not unusual for
the potassium concentration to equal or even exceed the
sodium concentration. This can be seen in analysis 9,
table 14, which represents water from sandstone, and
analyses 1 and 8, table 16, representing water from
dolomite and quartzite, respectively. Analysis 7, table
16, for the Wisconsin River, also shows this property.
Water containing 10 to 20 mg/L of sodium and
nearly equal concentrations of potassium appears to be
fairly common in the Ogalala Formation in southwestern
Nebraska (Johnson, 1960). However, in most other freshwater aquifers, if the sodium concentration substantially
exceeds 10 mg/L the potassium concentration commonly
is half or a tenth that of sodium. Concentrations of
potassium more than a few tens of milligrams per liter
are decidedly unusual except in water having high dissolved-solids concentration or in water from hot springs.
The concentrations of potassium in the two most saline
waters represented in table 17 (analyses 7 and 8) are
high, but potassium:sodium molar ratios are less than
0.02.
Significance

Seawater contains 390 mg/L of potassium (table


2). Analysis 6, table 14, representing mine water, and
analysis 3, table 18, representing an acidic thermal spring,
both have rather high potassium to sodium ratios.
There are deficiencies in the basic information about
the chemistry of potassium in water. Many of the available
chemical analyses of water and some of the analyses
included in this book do not include potassium determinations. Some of the old wet-chemical methods for potassium required considerable analytical skill, and the accuracy of some of the earlier published values for potassium
concentrations obtained using these methods appears to
be questionable. In more recent years, better methods
have become available, but potassium has remained one
of the more difficult ions to analyze accurately. Research
aimed at attaining a better understanding of the chemical
behavior of potassium in natural aqueous systems is
needed.
Alkalinity

and Acidity

The properties of alkalinity and acidity are important


characteristics of natural and polluted waters and are
almost always included in the chemical determinations.
However, these properties differ in important ways from
most of the other determinations reported in the analysis.
Both are defined as capacity functions-that
is, the
capacity of the solution to neutralize acid or base. Both
properties may be imparted by several different solute
species, and both are evaluated by acid-base titration, to
appropriate end points. Systems having these properties
are commonly referred to by chemists as buffered systems.
Most quantities determined in chemical analyses
are intensity functions-that
is, they are actual concentrations of a particular dissolved species at the time of
analysis. The measurement of pH provides values of
concentrations of H and OH- in solution. These species
contribute, of course, to acidity or alkalinity, but within
the pH range commonly seen in natural water they are
minor constituents, at concentrations of 10m.omolar or
lower. The principal solutes that constitute alkalinity are
imparted to natural water during its passage in liquid
form through the hydrologic cycle. They reflect the
history of the water, as an imprint left by these encounters.
The properties of alkalinity or acidity also evaluate the
potential of the solution for some kinds of water-rock
interaction or interaction with other material the water
may contact.
Most natural waters contain substantial amounts of
dissolved carbon dioxide species, which are the principal
source of alkalinity and can conveniently be evaluated
by acid titrations. Undissociated dissolved carbon dioxide
contributes to acidity rather than to titratable alkalinity
and can also be determined by titration using a basic
solution.
of Properties

and Constituents

Reported

in Water

Analyses

105

You might also like